0% found this document useful (0 votes)
32 views37 pages

ChemistryOpen - 2015 - Batchelor McAuley - Recent Advances in Voltammetry

The document reviews recent advancements in voltammetry, highlighting improvements in modeling, simulation, and various electrochemical techniques over the past decade. Key areas of focus include multistep electrochemical processes, voltammetry in ionic liquids, and nanoparticle electrochemistry, along with the theoretical frameworks of electron transfer. The review emphasizes the importance of accurate modeling and the application of advanced pulse techniques for better understanding complex electrochemical systems.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
32 views37 pages

ChemistryOpen - 2015 - Batchelor McAuley - Recent Advances in Voltammetry

The document reviews recent advancements in voltammetry, highlighting improvements in modeling, simulation, and various electrochemical techniques over the past decade. Key areas of focus include multistep electrochemical processes, voltammetry in ionic liquids, and nanoparticle electrochemistry, along with the theoretical frameworks of electron transfer. The review emphasizes the importance of accurate modeling and the application of advanced pulse techniques for better understanding complex electrochemical systems.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 37

DOI: 10.1002/open.

201500042

Recent Advances in Voltammetry


Christopher Batchelor-McAuley,[a] Enno K•telhçn,[a] Edward O. Barnes,[a]
Richard G. Compton,*[a] Eduardo Laborda,[b] and Angela Molina[b]

Recent progress in the theory and practice of voltammetry is ories of electron transfer (Butler–Volmer and Marcus–Hush), ad-
surveyed and evaluated. The transformation over the last vances in voltammetric pulse techniques, stochastic random
decade of the level of modelling and simulation of experi- walk models of diffusion, the influence of migration under con-
ments has realised major advances such that electrochemical ditions of low support, voltammetry at rough and porous elec-
techniques can be fully developed and applied to real chemi- trodes, and nanoparticle electrochemistry. The review of the
cal problems of distinct complexity. This review focuses on the latter field encompasses both the study of nanoparticle-modi-
topic areas of: multistep electrochemical processes, voltamme- fied electrodes, including stripping voltammetry and the new
try in ionic liquids, the development and interpretation of the- technique of ‘nano-impacts’.

Introduction

The last decade or so has realised remarkable progress in the linear diffusion problems, subsequently developed to more
area of voltammetry—the interrogation of an electrode reac- complex problems, but above all providing the impetus for
tion by means of exploring, under dynamic conditions, the cur- the physical electrochemist to model his or her own systems.[3]
rent-voltage characteristics of the process of interest so as to It is the rigorous comparison of theoretical and experimental
reveal, at one level, kinetic and mechanistic detail and, at an- results[3–4] that underpins the approach advocated in this
other, more fundamental level, the validity or otherwise of the- review, highlighting how such methods enable greater physical
ories of electron transfer. insights into the dynamics of complex systems and the under-
Voltammetric measurements are easily (and cheaply!) carried lying operative chemistry.
out, and data rapidly accumulated. However, interpretation of This article surveys the progress made, not only in simula-
the latter is often challenging, even to the well initiated espe- tion, but also in analytical theory, and comprises seven main
cially if quantitative information is sought. In particular, until subject areas of interest. First (Section 1), the use of the
relatively recently, data analysis required the use of analytical Butler–Volmer equation for modelling electrochemical reac-
equations confined by mathematical necessity to ‘model’ (or tions is considered, specifically in terms of its application to
‘toy’) systems under well-defined conditions of transport, elec- multistep processes; this area is then further developed
trode kinetics, and mechanism. As such, the area was often through consideration of the voltammetric response of electro-
limited to the study of experimental systems which were ame- active species in ionic liquids. Ionic liquids are of both theoreti-
nable to data analysis rather than being driven by the physico- cal and practical interest due to their large electrochemical
chemical interest of the system. The switch from this to real windows and the commonly observed altered chemical reac-
systems has been triggered by the ability to accurately simu- tivity. Second (Section 2), the physical validity of the Butler–
late the voltammetric problems dictated and driven by chemis- Volmer equation is questioned, and the development of the
try rather than constrained by what is theoretically possible. Marcus–Hush theory of electron transfer is probed from both
The origins of this essential switch lies in the pioneering work a theoretical and experimental standpoint. Importantly, the
of Rudolph[1] and the subsequent commercialisation of his soft- presented ‘asymmetric Marcus–Hush’ theory enables physical
ware in the form of the package DIGISIM.[2] This initially provid- reinterpretation of the Butler–Volmer equation which, in many
ed the first general basis of the modelling of ‘nonstandard’ cases, helps to validate its continued and historical use.
Recent voltammetric studies of the theories of electron
[a] Dr. C. Batchelor-McAuley, Dr. E. K•telhçn, Dr. E. O. Barnes, transfer have, in a number of cases, been aided and facilitated
Prof. R. G. Compton by the use of pulse techniques. Moreover, such pulse tech-
Department of Chemistry, Physical and Theoretical Chemistry Laboratory
niques are also of distinct importance in the broad area of
University of Oxford, South Parks Road, Oxford OX1 3QZ (UK)
E-mail: [email protected] electro-analysis. Consequently, Section 3 of this review is dedi-
[b] Dr. E. Laborda, Prof. A. Molina cated to the theory of these often complex but highly impor-
Departamento de Qu†mica F†sica, Facultad de Qu†mica tant techniques. In particular ‘differential double pulse voltam-
Regional Campus of International Excellence ‘Campus Mare Nostrum’ metry’, ‘additive differential pulse voltammetry’, ‘reverse pulse
Universidad de Murcia, 30100 Murcia (Spain)
voltammetry’, and ‘square-wave voltammetry’ are considered;
Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA.
a brief discussion of staircase cyclic voltammetry is also provid-
This is an open access article under the terms of the Creative Commons
Attribution License, which permits use, distribution and reproduction in ed. In combination with this work, the concept of the diffusion
any medium, provided the original work is properly cited. layer is advanced, enabling insight into the concentration pro-

ChemistryOpen 2015, 4, 224 – 260 224 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
file of electroactive species at the interface and diffusive mass Section 1: Advancing Butler–Volmer Theory
transport as a whole. Following from this, Sections 4 and 5
specifically continue in the development of ideas surrounding
1.1 Multistep electrode processes
mass transport. First, Section 4 reviews developments in the
use and modelling of diffusive random walks. Such theoretical Having performed a voltammetric experiment, observed a vol-
models are invaluable for the interpretation of newly develop- tammetric wave, and correlated its electrochemical presence
ing stochastic single-molecule and single-nanoparticle tech- to a given redox species in solution, the next common scientif-
niques, facilitating new insights into mechanisms on the nano- ic line of enquiry is to try and discern the operative electro-
scale. Second, Section 5 looks at the influence of migration chemical mechanism. A full review of this area of study is
upon the voltammetric response of electroactive species under beyond the scope of the current text, and the interested
conditions of low support. Beyond being a hindrance, low con- reader is directed towards the seminal lecture series by Sa-
centrations of supporting electrolyte can greatly facilitate elec- v¦ant[7] and the in depth reviews by Evans.[8] However, with
trochemical investigations, allowing mechanistic insights to be the recent publication of the IUPAC recommendations on the
gained from the differing charges of the intermediates and transfer coefficient, the area of multistep electrode processes
their subsequent interaction with the electric field, as de- warrants brief attention.
scribed by the Nernst–Planck–Poisson system of equations. The transfer coefficient (a) is now defined as;[9]
Having covered mass transport in relative detail, Section 6
considers the influence of altered electrode morphology upon  Œ Œ
RT d lnj jj
the voltammetric response, where the local surface structure a¼œ ð1Þ
F dE
(rough, porous, etc.) serves to alter the diffusion regime local
to the interface. The results of this section are of utmost im- where the flux j has been “corrected for any changes in the re-
portance for defining and evidencing authentic electrocatalysis. actant concentration at the electrode surface with respect to
Finally, in light of the above discussion, Section 7 turns to con- its bulk value”. The sign in Equation 1 depends upon whether
sider nanoparticle electrochemistry. The section discusses the reaction is anodic or cathodic. This IUPAC definition of the
nanoparticle-modified electrodes and the expanding field of transfer coefficient usefully and deliberately does not presup-
stochastic ‘nano-impact’ experiments. Systems in which the pose anything about the operative electrode reaction mecha-
nanoparticles mediate an electrochemical process are consid- nism. Mass-transport corrections to attain the flux j relative to
ered, as is the direct oxidation or reduction of the nanoparti- its bulk value is relatively facile for steady-state voltammetry.[10]
cles. However, for macroelectrode cyclic voltammetry, complete ex-
Although this review encompasses a very significant body of traction of this information is more involved, but, as demon-
the work available within the literature, the areas highlighted strated by Henstridge and Compton, certainly still obtaina-
are necessarily selective and focus upon fields deemed to be ble.[11] In most literature, experimental measurement of the
of particular contemporary importance; however, notable ab- transfer coefficient from macroelectrode voltammetry is limited
sences include both AC voltammetry[5] and hydrodynamic[6] to assessment of the Tafel slope at low overpotentials where
techniques, amongst others. the concentration of the reactant at the electrochemical inter-
face is not significantly altered from that of the bulk value.
Note classical literature tends to quote the magnitude of the
Richard G. Compton is Professor of Tafel slope (d E/d log j j j) which is directly related to reciprocal
Chemistry and Aldrichian Praelector at of the transfer coefficient. Having experimentally measured the
Oxford University, UK, where he is also transfer coefficient, its interpretation provides one route by
a Tutor in Chemistry at St. John’s Col- which the electrode mechanism may be elucidated, as will be
lege. His research covers fundamental discussed below.
and applied electrochemistry and elec-
For a one-electron redox process with unit stoichiometry as
troanalysis. He has published over
given by
1300 papers (h = 82; cites, excluding
self-cites > 26 000; Web of Science,
Thomson Reuters, March 2015). The A œ e¢ Ð B ð2Þ
2nd edition of his textbook Understand-
ing Voltammetry (with Craig E. Banks) under reversible (quasi-equilibrium1) conditions the transfer co-
was published in 2011 (Imperial College Press). He is a Lifelong efficient simply reflects the number of electrons transferred,
Honorary Professor at Sichuan University, China, and is the Found- which in this case is one. This value for the reversible case is
ing Editor and Editor-in-Chief of the journal Electrochemistry Com- 1
Although under reversible conditions, the surface concentrations of the elec-
munications (current Impact Factor = 4.2, Journal Citation Reports troactive species are defined by the Nernst equation, and hence may be
2014, Thomson Reuters) published by Elsevier. He is also an Editori- viewed as being at ‘equilibrium’; the mass-transport regime is time variant
al Advisory Board member of ChemistryOpen. and hence the term ‘equilibrium’ is caveated by the prefix ‘quasi’. This should
not be conflated with the term ‘quasi-reversible’ which implies the system’s
deviation away from equilibrium at the electrochemical interface.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 225 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
generally referred to as an ‘apparent transfer coefficient’. The have a profound influence.[15] Experimentally, this situation is
term ‘apparent’ is used to emphasise that the measured trans- well exemplified by quinone reductions. In non-aqueous sys-
fer coefficient does not reflect the underlying electron transfer tems two one-electron waves are observed; the addition of
kinetics, but originates directly from the system being under water serves to compress the potential difference between the
Nernstian control. Under irreversible conditions, the transfer two waves until they merge to yield one voltammetric wave.
coefficient takes a value between 0 and 1, but is commonly However, in aqueous media (at high pH), it has been experi-
0.5 œ 0.2 for a one-electron process. As a note of caution, the mentally demonstrated that the potential of the first and
transfer coefficient as measured from a voltammogram may be second electron transfers can be ‘tuned’ through ion-pairing.[16]
‘artificially’ distorted due to a potentiostats application of stair- A consequence of the first and second electron transfer occur-
case as opposed to a true analogue voltage ramp; this point is ring at similar potentials is that, although only one voltammet-
discussed in further detail in Section 3.4. ric wave is observed, the peak height is suppressed, and the
Use of the Butler–Volmer equation to model an electro- wave is broader than would be anticipated if the electrons
chemical system implicitly assumes the transfer coefficient to were assumed to be transferred simultaneously.[17] In such
be constant as a function of potential; deviations from such cases, the use of square-wave voltammetry is advisable for the
linear Tafel behavior have been experimentally evidenced. precise measurement of the associated formal potentials.[18]
Moreover, the Butler–Volmer equation is periodically damned The presence of coupled homogeneous chemical processes
due to being ‘phenomenological’ in its description of electron- or structural changes in the molecular structure may lead to
transfer, that is, it has no direct physical meaning. However, as a situation known as potential ‘inversion’ where the second
will be expanded upon in Section 2 of this review, the Butler– electron transfer is easier than the first, that is (Ef1¢Ef2) is nega-
Volmer equation can be better understood as often being tive.[19] Reportedly, a potential inversion of 400 mV is required
a highly precise approximation of electron-transfer rates at low such that ‘no’ intermediate is formed.[20] Figure 1 depicts the si-
overpotentials2 and warrants its use due to its mathematical mulated voltammetric response for a two-electron reduction
simplicity. Moreover, comparison of the Butler–Volmer theory as described by the mechanism above. Also shown is the pre-
with asymmetric Marcus–Hush theory allows a physical under- dicted voltammetric response for the hypothetical situation in
standing of the transfer coefficient (see Section 2 for further which two electrons are transferred simultaneously. This result
details) in terms of the changing force constants in the redox clearly demonstrates how even for this simplest multistep pro-
reaction. Before moving on any further, it is highlighted that cess, the corresponding voltammetric wave shape can vary sig-
the terms ‘reversible’ and ‘irreversible’ when applied to voltam- nificantly when the two respective formal potentials are com-
metry refer to the magnitude of the electron transfer rate rela- parable in magnitude, a situation which is commonly encoun-
tive to the prevailing rate of mass transport to and from the tered. Two important insights should be taken from this: first,
electrochemical interface.[12] This is an important point that tac- a potential inversion of greater than 100 mV is required for the
itly underpins much of the work discussed within this body of peak current of the stepwise two-electron voltammetric re-
text. It is also a general lack of insight into this problem that sponse to be within 3 % (i.e. within experimental error) of that
regularly leads to misinterpreted and misreported results obtained for the ‘simultaneous’ case. Second, a common route
within the literature (see Section 7 for further discussion). to determining the diffusion coefficient of an analyte is to
One of the principal points in the IUPAC recommendation is
that the simultaneous transfer of two or more electrons is
highly unlikely. Consequently, mechanistic interpretation of the
transfer coefficient must be done in the light of this fact![9] At
this stage, the simplest multistep electrochemical process is
considered: that of two sequential reversible one-electron
transfers (an ErevErev reaction in Testa and Reinmuth notation[13])
as described by

A œ e¢ Ð B E f1 ð3Þ

A œ e¢ Ð C E f2 ð4Þ

In the absence of coupled homogeneous kinetics the


second electron transfer is likely to be less favorable purely
due to coulombic repulsion.[8] However, the relative potentials
at which these two redox processes occur is highly solvent Figure 1. Simulated voltammetric response of a two-electron transfer as
and electrolyte dependent;[14] notably hydrogen bonding can a function of the difference between the formal potentials for the two pro-
cesses. Ef1¢Ef2 : + 50 mV (magenta), 0 mV (blue), and ¢50 mV (red). The black
2
Note that for many experimental cases, the potentials at which significant de- line represents the hypothetical limiting case in which the two electrons are
viations from Butler–Volmer kinetics occur are not easily experimentally ac- transferred simultaneously. Simulations performed using DigiSim; other sim-
cessible due to the reaction coming under mass-transport control. ulation parameters: v = 100 mV s¢1, D = 10¢5 cm2 s¢1, r0 = 1 mm.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 226 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
measure the respective peak current as a function of scan rate, and outdated as it implies the possibility of multiple electrons
where the peak current is proportional to the square root of being transferred simultaneously. A problem arises upon rec-
the diffusion coefficient. In the case of a multistep process, as ognition that many older analytical expressions used for inter-
highlighted by Figure 1, this may lead to a significant error preting voltammograms, which are still regularly used, are in
due to the sensitivity of the peak current to the formal poten- error—one prime example being the use of the variation of
tials of the electron transfers.[21] the peak potential of a surface-bound redox wave as a function
As a secondary example, the case where an electron transfer of scan rate to extract kinetic and mechanistic information.[25]
is coupled to a homogeneous reaction as given by the mecha- In this model, the peak position is related to an; hence, use of
nism: such expressions for analysis of the operative mechanism will
be in error.
A œ e¢ Ð B ð5Þ As commented above in reference to electrochemical rever-
sibility, voltammetry is inherently the study of interfacial pro-
BÐC ð6Þ cesses; consequently, understanding of a system’s response re-
quires an understanding of the prevailing mass-transport
is considered. This scheme is known as an ‘EC reaction’ where regime and its effect. An example of this for multistep process-
E signifies an interfacial redox process and C is a coupled ho- es relates to the influence of an electrode’s size. On decreasing
mogeneous reaction. For the case in which the electrochemical the dimensions of an electrode, that is from macro to micro,
step is reversible, as the chemical step C is made more thermo- the mass transport to and from the surface becomes more effi-
dynamically favorable (assuming it is not kinetically hindered), cient.[26] This is akin to increasing the rotation rate of a rotating
the potential required for the redox process decreases in mag- disk electrode, however far higher mass-transport rates are at-
nitude, with a corresponding loss of the voltammetric back tainable with the use of microelectrodes.[27] The decrease in
peak. As the chemical step becomes more highly driven, the the electrode size results in the electron transfer process be-
redox wave shifts to a potential at which the rate of electron coming more irreversible and increases the probability of inter-
transfer becomes the rate-determining step (note the rate of mediates being released. To more succinctly restate this, on
electron transfer decreases exponentially, in accordance with changing the electrode size the overall electrochemical mecha-
the Butler–Volmer equation). Hence a ‘reversible’ electron nism may be altered, resulting in the possible release of
transfer process (i.e. one which has a high k0) may appear elec- ‘higher’-energy intermediates. One example of such a situation
trochemically irreversible when the product is consumed by is found with oxygen reduction at silver surfaces. At a macroe-
a highly driven chemical step.[22] lectrode the process is found to, on average, involve the trans-
It can be seen from the above two examples that the com- fer of 3.3 electrons to each oxygen molecule.[28] This corre-
plexity of an electron transfer reaction rapidly increases with sponds to roughly one in three oxygen molecules undergoing
the number of steps, and hence such systems are best under- only a two-electron reduction to hydrogen peroxide before re-
stood through simulation (further discussion of multistep elec- lease from the electrode, as opposed to the full four-electron
trochemical processes can be found in the work of Batchelor– reduction. However, decrease of the size of the electrode leads
McAuley and Compton[21]). This is especially true for situations to an increase in this ratio, such that the probability of hydro-
involving proton transfer where a multitude of possible mech- gen peroxide production is increased, where on the nanoscale,
anistic routes are possible, and the exact pathway (or path- almost all formed hydrogen peroxide is released prior to fur-
ways) taken will depend strongly upon the pH, the electrode ther reduction.[28] This insight that reactivity may change on
potential, the pKa value associated with the reactants, inter- the nanoscale solely due to the altered mass transport (i.e.
mediates, and products, and the electron-transfer kinetics of even without considering plausibly altered nanoparticle ther-
individual steps.[17] Related to this is the investigation of sys- modynamics or the expression of higher-order crystal facets)
tems in which the proton and electron are transferred simulta- has two implications: first, it implies a limitation on the effi-
neously, a reaction which is possibly of distinct importance in ciency of the use of nanoparticles for catalysis in industrially
biology.[23] relevant processes.[25] Secondly, and perhaps more importantly,
Although full understanding of an electrochemical process is the changed reactivity on the nanoscale arising from the al-
best achieved through simulation, the rate-determining step of tered mass-transport regime may have wide-ranging implica-
a multistep mechanism may be readily assessed by experimen- tions for the nanotoxicity of these materials towards biological
tal measurement of the Tafel slope, noting that the magnitude systems.[29] It is this changed reactivity at diffusionally isolated
of the Tafel slope may also vary as a function of scan rate due particles that highlights one of the driving forces for wishing
to the presence and influence of coupled homogeneous kinet- to study reactions at individual nanoparticles, a subject that
ics.[24] The Tafel slope should be interpreted as being equal to will be focused upon more within the final section of this
¢(n’ + aRDS)F/RT (for a cathodic process) where n’ is the review.
number of electrons transferred before the rate-determining The following section focusses on the simulation and under-
step, and aRDS is the transfer coefficient of the rate-determining standing of voltammetry in ionic liquids, a medium in which
electron transfer.[21] Noting that again, in accordance with the mass-transport rates are appreciably slower than found for
IUPAC, assuming that the Tafel slope is equal to ¢(an)F/RT, common aqueous systems.
where n is the total number of electrons transferred is invalid

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 227 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1.2 Voltammetry in ionic liquids A œ ne¢ Ð B ð7Þ

The last decade has seen an explosion of electrochemical inter- independently of the voltammetry, ideally using potential-step
est in the use of room-temperature ionic liquids (RTILs): liquids or double-step chronoamperometry.[42]
composed entirely of ions which only solidify at temperatures A third but related issue is that the diffusion coefficients of
well below ambient.[30] This is partly because they offer signifi- the analytes in RTIL media can be quite sensitive to the pres-
cant advantages for some applications, most notably in energy ence of dissolved gases not least because of the high solubili-
transformation technology[31] and in gas sensing,[32] but also ties (~ m) of species such as H2S or CO2 in many RTILs, but also
since they challenge existing theories of electron transfer[33] because the dissolved gases significantly perturb the solvent
and interfacial structure.[34] Progress in the area has been regu- structure and hence the transport of the other solutes.[43]
larly reviewed.[30, 35] This section focuses on the altered voltam- A fourth issue relating to the different voltammetry in RTIL
metry seen in RTIL media. media as compared to molecular solvents is that the common
From an electrochemical perspective, RTILs offer some im- approximation of assuming equal diffusion coefficients for
portant contrasts with conventional solvents such as water, most or all species involved in an electrode reaction is usually
acetonitrile, THF, etc.. First, the potential window displayed, de- adequate for the quantitative simulation of the voltammetry;
fined by the onset of cathodic and anodic solvent decomposi- this approximation holds much less well for RTIL media. The
tion, is unusually wide and, for rigorously dried solvents, can reason for this is that because of the ionic nature of the sol-
extend to as much as 5 or 6 V.[36] This reflects the stability to- vent, the transport properties are sensitive not only to the
wards oxidation and reduction of the component ions which solute size (Stokes-Einstein equation) but also to the solute
are generally a bulky organic cation and a small inorganic charge.[44] One extreme example is the one-electron reduction
anion such that the size mismatch discourages crystallisation of oxygen,
except at unusually low temperatures. Indeed the use of the
same materials as supporting electrolyte in conventional sol- O 2 þ e¢ Ð O 2 C ¢ ð8Þ
vents has been advocated, reflecting the intrinsic inertness of
the ions.[37] in the RTIL hexyltriethylammonium bis(trifluoromethyl)sulfonyl
A second significant difference lies in the observation[30] that imide, where at 25 8C
many ionic liquids have viscosities which are larger—often an
order of magnitude greater—than conventional molecular or-
DO2 ¼ 1:5 10¢10 m2 s¢1
ganic solvents. This is reflected in the magnitude of the diffu-
sion coefficient of the solutes which, except for rather small
DO¡2 ¢ ¼ 4:7 10¢12 m2 s¢1
sized molecules such as O2[38] or H2S,[39] generally reflect the
Stokes–Einstein equation, with diffusion coefficients scaling in-
which gives a ratio of diffusion coefficient of over 30![45] The
versely with the viscosity.[40] The much reduced rates of diffu-
consequence of this marked difference is that the voltammetry
sion in RTIL media have the very important consequence that
leads to curious current-voltage response in which a microdisc
the transition from linear to fully convergent diffusion as ob-
electrode steady-state voltammogram is seen for the forward
served at microelectrodes occurs at quite different (much
scan, corresponding to the faster diffusing O2 reduction,
lower) voltage scan rates than are familiar to electrochemists
whereas in the reverse scan a peak is observed for the reoxida-
operating in aqueous or non-aqueous media.[41] As a result, it
tion of O2C¢ because the slowness of its diffusion leads to the
is quite common to see peak-shaped rather than sigmoidal
accumulation of O2C¢ near the electrode surface, and the trans-
current voltage curves when using microelectrodes in ionic
port contains a significant component of linear diffusion.[45]
liquid media, and true steady-state diffusion limited currents
can be difficult to observe unless unusually if not pathological- The discussion above has focused on measurements made
ly slow scan rates are deployed (under which conditions other in RTIL media using microelectrodes, which are a preferred
factors such as slow adsorption or coupled kinetics may unde- methodology for such investigations.[32, 46] This is because the
sirably kick in). It follows that the extraction of kinetic and deployment of microelectrodes facilitates the use of small vol-
transport parameters from the voltammetry requires the nu- umes (~ 10 mL) of solvent, which is important in the RTIL area
merical simulation[41] of the voltammetry rather than the appli- since it is vital to properly dry the solvent and to ensure that
cation of the simple analytical equations derived for pure they are water free. The drying of RTILs is readily undertaken
linear diffusion (Randles–Ševč†k equations) and for pure con- using a T-cell arrangement[32, 46–47] in which the solvent can be
vergent diffusion at a microdisc (I = 4nFDCr, where n is the exposed to vacuum and rigorously dried before voltammetric
number of electrons transferred, F is the Faraday constant, D is study. Any residual water will greatly reduce the electrochemi-
the diffusion coefficient, C is the analyte concentration, and r is cal window[36b] and markedly alter the diffusion coefficients of
the radius of the electrode). Moreover, the fitting of such simu- the solutes.[46] The drying of larger quantities of solvent is
lations is challenging, but can be helpfully simplified if the dif- a slow process because of the need for the water to diffuse to
fusion coefficients of the reactant, A, and product, B, are deter- the liquid surface and evaporate; as observed diffusion in RTILs
mined for: can be a very slow process, and so the purification of the
much larger volumes required for macroelectrode experiments

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 228 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
would be time-consuming and possibly incomplete. Indeed possible. The following section looks at recent advances in the
the lack of reproducibility of simple data such as diffusion co- development of models of electron transfer kinetics.
efficients (even of ‘model’ compounds such as ferrocene), in
early work in the field probably reflects the different composi-
tion of the solvent used in terms of dissolved water and gases.
Finally, we consider whether Butler–Volmer theory is applica- Section 2: Challenging Butler–Volmer Theory
ble in RTIL media noting the near-ubiquitous success claimed In recent years, a renewed interest in the suitability of the
for the phenomenological approach in molecular solvents. The available kinetic models for heterogeneous electron transfer re-
first consideration is to note that the slowed diffusion in RTILs actions has led to critical assessment[61] of the most well-estab-
promotes the apparent electrochemical reversibility of many lished approaches: the Butler–Volmer (BV)[62] and the Marcus–
redox couples. In order for the voltammetry to reveal quasi- or Hush (MH)[63] models.
irreversible electrode kinetic behavior, it is required that the The BV formalism has been preferred over the years (and
studied rate constant must fulfill k0 < mT, where mT (cm s¢1) is still in the present) due to its simplicity and satisfactory de-
the mass-transport coefficient of the electrode. Typically this is scription of the electrode kinetics of many systems with three
approximated by mT ~ D/r, where r is the electrode radius, so fitting parameters: the standard heterogeneous rate constant
that even with microelectrodes it can be challenging to extract (k0), the transfer coefficient (a) and the formal potential (Ef0 ).
electrochemical rate constants from the voltammetric data. It Thus, the rate constants are given by the following well-known
follows that measurements made using macroelectrodes in expressions:
RTIL are unlikely to give reliable data especially since it is
noted that the conductivity of many ionic liquids is similar to Š ¨ ¦‰
F E ¢ Ef0
that of conventional organic solvents (DMF, CH3CN, THF, etc.)
BV
kred ¼ k0 exp ¢a ð9Þ
RT
containing about 0.1 m supporting electrolyte so that macro-
electrode voltammetry in RTILs is also typically as distorted by Š ¨ ¦‰
F E ¢ Ef0
ohmic losses as is voltammetry in organic media. BV
kox ¼ k 0 exp ð1 ¢ aÞ ð10Þ
RT
Extensive modeling of a wide diversity of voltammetric sys-
tems has been undertaken using small microelectrodes in
In spite of being successful in kinetic parameterisation of
order to provide a better possibility of extracting kinetic pa-
a great majority of redox systems, easy-to-implement, and
rameters. Systems studied include O2/O2C¢ ,[45] Br¢/Br2,[48] nitro-
computationally inexpensive, the BV expressions for the rate
benzenes,[49] aryl amines,[50] NO2/NO2¢/NO2 + ,[51] I¢/I2,[52] aromatic
constants are empirical (but see below in connection with the
diamines,[53] arenes,[54] Li/Li + ,[55] benzoquinone,[56] hydroqui-
discussion of asymmetric Marcus–Hush theory). Therefore, the
none,[57] and H + /H2.[58] In many cases the values of k0 obtained
adjustable parameters provide limited physical insight in terms
correspond to quasi-reversible behaviour and, as such, do not
of the nature of the electroactive molecules, the medium, and
provide a perfect test of the validity of Butler–Volmer kinetics
the electrode, and it is not possible to make predictions. More-
since the behaviour is approximately Nernstian. To restate this,
over, experimental deviations from the ever-increasing expo-
under reversible (Nernstian) conditions, no information regard-
nential variation of the rate constants with E ¢ Ef0 predicted by
ing the kinetics of the electron transfer process may be infer-
Equations 9 and 10 have been reported.[64]
red from a voltammetric experiment. Hence, as a system tends
towards reversibility, obtaining unambiguous results evidenc- The above limitations of BV calls for the use of more realistic
ing the validity or otherwise of the applicability of the Butler– models that enable us to fully describe the experimental data,
Volmer equation becomes inherently more challenging. How- as well as to connect the electron transfer kinetics with the
ever, in some cases, notably the I¢/I2 system,[52] the oxidation nature of the system. With this aim, in recent years the applica-
of hydroquinone,[57] Li/Li + ,[55] and the H + /H2 system[58] there is bility of the Marcus–Hush model has been theoretically and ex-
clear electrochemical irreversibility, and the accuracy of Butler– perimentally assessed via voltammetry in its symmetric[64b] and
Volmer kinetics in reproducing observed experimental behav- asymmetric[65] versions.
ior is excellent. Note that the follow-up chemistry in these sys-
tems ‘promotes’ the irreversibility of the system and, as such,
multistep processes may be preferred for studying electron
2.1 The symmetric Marcus–Hush model (sMH)
transfer in RTILs.[59]
Very recently,[60] an attempt has been made to apply The symmetric version of Marcus theory[66] considers the pa-
Marcus–Hush theory to ionic liquids focusing on the O2/O2 .¢ rabolas describing the Gibbs energy of the reactants and prod-
couple. Solvent reorganisation energies around 0.4–0.5 eV ucts to be of equal curvature. The heterogeneous electrochem-
were found and attributed to inner sphere reorganisation with ical reaction between a molecule and a metallic electrode in-
a negligible contribution from solvent reorganisation. volves transfer of charge from a discrete molecular energy
In summary, quantitative voltammetry in RTILs present spe- level to a continuum of states (e), with the energy levels in the
cial challenges, but with the aid of numerical simulation, quan- electrode occupied according to the Fermi–Dirac distribution.
titative understanding of the both kinetics and mechanism is This model leads to the following expressions for the hetero-
geneous rate constants:[64b]

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 229 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
MH Sred ðh; LÞ
kred ¼ k0 ð11Þ
Sred ð0; LÞ

MH Sox ðh; LÞ
kox ¼ k0 ð12Þ
Sox ð0; LÞ

where Sred=ox ðh; LÞ are integrals:

h i
Z y
1 exp ¢DGsym;red=ox ð x Þ=RT
Sred=ox ðh; LÞ ¼ dx ð13Þ
¢1 1 þ expðm › x Þ

y
with DGsym;red=ox ð x Þ being the activation energy of the electro-
reduction/oxidation process for each electronic level that ac- Figure 2. Schematic of the parabolic Gibbs energy curves as given by the
cording to the symmetric version of the Marcus theory is given asymmetric version of the Marcus theory.
by:

y that, within the asymmetric Marcus theory, can be written


DGsym;red=ox ð x Þ L  h þ x Ž2 ð14Þ
¼ 1œ as:[65a, 66]
RT 4 L
F ¨ ¦ F
where h ¼ RT E ¢ Ef0 , x ¼ RT ðe ¢ E Þ, and L is the dimension- y Š h þ xŽ2 ‰
F DGasym;red=ox ð x Þ L  h þ xŽ2 h þ xŽ L
less reorganisation energy: L ¼ RT l. When two signs appear, ¼ 1œ þg 1¢ þ g2
RT 4 L L L 16
the upper sign refers to reduction and the lower sign refers to
oxidation. For the calculation of the integrals of the sMH for- ð15Þ
malism [Eq. (13)], numerical integration methods can be em-
ployed[67] and analytical approximations have also been pro- with the parameter g being defined as:
posed to make the implementation of the sMH expressions
easier.[68] P ¨ ¦2
li l s ks Dq0 ls
As with the BV model, sMH theory describes the electron g¼ hls i ¼ i P ¨ s ¦2 ð16Þ
l l s ks Dqs
0
transfer kinetics as a function of three adjustable parameters:
the formal potential, the standard heterogeneous rate con-
where li is the inner-sphere reorganisation energy, Dq0s is the
stant, and the reorganisation energy (l). The latter corresponds
difference between the equilibrium values for the s-th normal
to the energy required to distort the atomic configurations of
mode coordinate of reactant and product, and ks and ls are
the reactant molecule (inner-sphere component of l) and its
symmetric and antisymmetric combinations of the force con-
solvation shell (outer-sphere component) to those of the prod-
stants of the s-th mode of the oxidised (fsox ) and reduced (fsred )
uct in its equilibrium configuration. Therefore, l enables us to
species:
rationalise the electrode kinetics in terms of the microscopic
nature of the system such that the larger the structural and
2fsred fsox
solvation changes as a consequence of the electron transfer, ks ¼ ð17Þ
fsred þ fsox
the larger the l value and the slower the electrode kinetics.

fsox ¢ fsred
ls ¼ ð18Þ
fsox þ fsred
2.3 The asymmetric Marcus–Hush model (aMH)
In the sMH model, the Gibbs energy parabolas are assumed to The values of the rate constants in the aMH model are calcu-
have the same curvature, which means that intramolecular vi- lated from the expressions presented in Section 2.1 by substi-
brations and solvation are, on average, the same for the re- tuting Equation 15 into 13. In the numerical integration of
duced and oxidised species. This may not hold as a general Equation 13, the limits of the integral must be restricted to the
rule given the different charge of the reduced and oxidised x-range where the integrand value is significant, typically œ 50.
species, and various theoretical approaches have been consid-
ered to overcome this limitation of the sMH formalism. Among From the definitions in Equations 16, 17, and 18, it is clear
them, the use of the asymmetric version of the Marcus theory that the g value relates to differences between the vibrational
has been recently applied to heterogeneous electron transfer force constants of the electroactive species such that it takes
processes by Compton et al.[61a] a positive value when the force constants of the oxidised spe-
As can be observed in Figure 2, different (vibrational and/or cies are greater (on average), a negative value in the opposite
solvation) force constants result in Gibbs energy curves of dif- situation, and g = 0 when the (average) force constants are
ferent curvature and affect the value of the activation energy equal. Note that the last particular case coincide with the sym-

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 230 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
metric MH model such that Equation 15 simplifies to 14 for
g = 0.
In the derivation of Equation 15, only the first terms in the
expansion of hls i have been considered so that the aMH for-
malism above presented accounts for differences between the
vibrational modes of the reduced and oxidised species with
only one additional fitting parameter with respect to BV and
sMH (g, [Eq. (16)]). Higher order terms in hls i[66] would be neces-
sary when the force constants differ significantly (by a factor of
more than 2[61a]), which would make the model less general
and more complex.

2.4 Voltammetric assessment of the BV and MH models


As shown in Figure 3 the different kinetic models predict differ-
ent variations of the rate constants with the applied potential.
Thus, whereas the reductive rate constant increases exponen-
tially and continuously as E ¢ Ef0 is more negative,
” in” the MH
models kred shows a limiting value at large ”E ¢ Ef0 ” values,
which is consistent with the curved Tafel plots and potential-
dependent transfer coefficients reported in the literature.[64]
The divergence from the BV behaviour is more apparent for
small l-values and at large overpotentials (Figure 3 b).
Another key point to consider in Figure 3 is that, independ-
ently of the l-value, the curves for the reduction and oxidation
rate constants are symmetrical with respect to the axis
E ¢ Ef0 = 0 in sMH, such that kred(E ¢ Ef0 ) = kox(¢(E ¢ Ef0 ). This
reciprocity relation[68a] breaks down in the aMH model when
g¼6 0. Thus, when the force constants of the oxidised species
are greater, g > 0, the cathodic branch is steeper than the
anodic one, and the opposite is true for g < 0. Note that the g-
effect is more significant at large overpotentials and it is analo-
gous to the effect of a in BV. Indeed, at low overpotentials, the
kred/ox values calculated from the aMH formalism tend to those
obtained in BV with the following transfer coefficient:
 Œ
¨ ¦ 1 1 1:267
a Ef0 ¼ þ g ¢ ð19Þ
2 4 L þ 3:353

Thus, the case a < 0.5 relates to force constants of the re-
duced species greater than those of the oxidised one (i.e., g <
0) and the opposite applies for a > 0.5. This enables physical
reinterpretation of the a data available from Butler–Volmer
analysis, extending over many years.
The effect of the asymmetric parameter g on the voltammet-
ric response is also analogous to that of a in BV. This is shown Figure 3. Variation of the reduction and oxidation rate constants with E¢Ef0
in the Butler—Volmer model (A), the Marcus—Hush models (B), and the
in Figure 4 for the response of diffusional systems in cyclic vol-
asymmetric Marcus–Hush model (l = 2 eV) (C).
tammetry and reverse scan square wave voltammetry under
transient conditions. In the latter, as well as in the reverse scan
of cyclic SWV, a cathodic peak and an anodic one can be ob-
served on either side of the formal potential in the case of values, as occurs in BV for a > 0.5. This fact points out the
sluggish electron transfers.[69] In both CV and SWV, the reorgan- greater flexibility of the aMH model for quantitative fitting of
isation energy affects the reductive and oxidative peaks of the the voltammetry through the new kinetic parameter.
voltammograms similarly (Figure 4 b), whereas the g-value has In summary, the kinetic formalisms discussed above predict
an effect on the relative anodic/cathodic peak heights and the different dependence of the reduction and oxidation rate con-
peak potentials. The reductive peak increases in height and stants with the applied potential, the divergence between
shifts to less negative potentials as g takes more positive them being more apparent for small values of the reorganisa-

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 231 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
other possible background pro-
cesses can be reduced. Differen-
tial pulse techniques in reverse
or cyclic modes are of particular
value given that, as shown in
Figure 4, they enable simultane-
ous examination of the reduc-
tion and oxidation processes,
which is essential to confirm the
consistency of the kinetic param-
eters obtained from the fitting
of experimental data.[61e] Also,
the value of the large amplitude
Fourier-transformed AC voltam-
metry has been examined.[70b]
The analysis of the higher order
harmonic responses and the fre-
quency-dependence of the peak
heights of the harmonics is pre-
dicted to be very powerful and
sensitive in the study of the ap-
plicability of the different kinetic
models and the extraction of ki-
netic parameters.
It is also worth highlighting
that only the aMH model is com-
patible with the asymmetric,
curved Tafel plots obtained ex-
perimentally for surface-bound
and diffusional redox systems,[64]
as demonstrated in work by
Henstridge et al.[65b] by the fit-
ting of experimental data avail-
able in the literature.[64c] Other
contrasting behaviours between
the kinetic formalisms have been
theoretically described and they
potentially allow for critical eval-
uation of the models, though
the experimental conditions nec-
essary are challenging. Thus, the
sMH and aMH models predict
Figure 4. Influence of the kinetic parameters of the different kinetic models on the response of diffusional quasi-
reversible and irreversible systems in cyclic voltammetry and reverse scan square wave voltammetry.
deviations from the Randles–
Ševč†k behaviour for irreversible
processes.[72] The experimental
tion energy and at large overpotentials. In order to point out evidence of such deviations is not straightforward, particularly
such differences experimentally and assess the suitability of in the case of diffusional systems, given that it requires the
the different kinetic models, various electrochemical methods study of systems with small reorganisation energy (unlikely in
have been proposed and employed in the literature as an al- the case of slow kinetics) in a broad range of scan rates.
ternative (or complement) to cyclic voltammetry.[70] The use of Another striking difference between the BV and the MH
differential pulse voltammetries (namely, square wave voltam- models is that in the latter, the limiting current (in single and
metry and differential multipulse voltammetry) has proven double-step chronoamperometry, as well as at fast-flow chan-
very insightful in revealing differences between the kinetic nel electrodes[70a]) can be smaller than the mass-transport-con-
models[71] as well as being very adequate for quantitative stud- trolled limit and depend upon the electrode kinetics.[61, 70a]
ies. Thus, due to the subtractive nature of these techniques, Again, this phenomenon is predicted to be more apparent for
well-defined, peak-shaped responses are obtained and undesir- small values of the reorganisation energy. Given that this is in
able distortions associated with double layer charging and general associated with fast kinetics (i.e., large k0 values), spe-

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 232 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
cial attention has been paid to the use of nanosize (including Section 3: Advances in Voltammetric Tech-
nanodiscs[61d, e] and impacting nanoparticles[73]), nanogap,[61b, c] niques
and channel[74] electrodes such that the enhanced mass trans-
port shifts the kinetic-controlled voltammetric response away
3.1 Double potential pulse techniques at microelectrodes
from Ef0 . Thus, it is theoretically possible to observe kinetically-
limited steady-state currents at large overpotentials in the In recent years the use of double potential pulse techniques
above systems when the size of the electrode or the gap dis- for the study of electrode kinetics and reaction mechanisms
tance is reduced to the nanometer scale, though in practice has been developed both theoretically and experimentally at
this requires that the geometry of the electrode is accurately microelectrodes.[77] The combination of pulse techniques and
known and, in the case of electrodes of a few nanometers, to small-sized electrodes offers important advantages in terms of
deal with double layer and nonclassical effects. accuracy as a result of the reduction of distorting effects
(mainly ohmic drop and charging current),[26, 78] which leads to
well-defined signals adequate for electrochemical studies even
2.5 Experimental assessment of the kinetic models
in media of low conductivity. With respect to electrode reac-
A critical study of the models presented above has recently tions complicated by coupled (electro)chemical processes
been undertaken by studying the voltammetric response of (Figure 5), analytical theory for double pulse techniques at mi-
various solution-phase systems, including the one-electron re- croelectrodes of different geometries has been developed for
ductions of 2-methyl-2-nitropropane, 1-nitropentane, 3-nitro- the study of the (pseudo)first-order CE,[79] EC,[79–80] catalytic[81]
phenolate, cyclooctatetraene, and europium(III), as well as the and equilibrium square[82] mechanisms as well as multistep
electro-oxidation of tetraphenylethylene.[61a, 65c] As concluded electrode processes.[83] Analytical expressions for one-electron
from Figure 3, in order to observe differences between the BV, transfer processes of solution-phase redox systems of any re-
sMH, and aMH models, the experimental system must give a ki- versibility degree have also been deduced for double potential
netically controlled current at appreciable overpotentials. This pulse techniques at (hemi)spherical microelectrodes. The rever-
has been achieved by Compton et al. for electrode processes sibility criteria and methodologies for kinetic analysis are ap-
with k0 ‹ 0.02 cm s¢1 by using microelectrodes of 25–50 mm propriate for other microelectrode geometries and will be dis-
radius, which also allows for the reduction of undesirable cussed in the following sections.
ohmic drop and capacitive effects. The use of nanosize[61d, e, 73]
or nanogap[61b] electrodes would be necessary for faster elec-
3.1.1 Differential double pulse voltammetry (DDPV) and addi-
tron transfers, which presents difficulties in terms of electrode
tive differential pulse voltammetry (ADPV)
fabrication and characterisation as well as modeling of non-
conventional effects. The subtractive nature of the DDPV and ADPV techniques (in-
The voltammetric response in different techniques (mainly troduced in the work of Molina et al.[84]) make them very valua-
cyclic and square wave voltammetries) of several one-electron ble for quantitative analysis since the influence of background
transfer processes without chemical complications and under currents is further reduced, and peak-shaped responses are ob-
fully-supported conditions has been analyzed making use of tained. The influence of the electrode kinetics on the DDPV
the three kinetic models. The sMH model has been unable to and ADPV signals are shown in Figure 6. The single-peak DDPV
fit the experimental voltammetry of systems with transfer coef- response and the double-peak ADPV signal shift to higher
ficients notably different from 0.5 (as those chosen in the ex- overpotentials, and the peaks become smaller and broader
perimental studies), which is expected in light of the results (larger half-peak width) as the electrode reaction transitions
discussed in Section 2.4. On the other hand, the BV and aMH between the fully-reversible and the fully-irreversible limits. In
formalisms yield satisfactory fittings of similar quality, with the the latter, the shape of the DDPV and ADPV signals is inde-
correlation between the parameters g and a above-mentioned pendent of the k0-value whereas the position does depend on
being found experimentally, such that a-values different from k0. Regarding the influence of the transfer coefficient (a), the
0.5 may be interpreted as an indicator of different force con- peak width increases, the peak height decreases, and the peak
stants in the oxidised and reduced species. Note that such dif- potential takes more negative values as the a-value is smaller
ferences can also arise from the interactions with the solvent in the case of electroreduction processes. The splitting of the
as theoretically demonstrated in Laborda et al.[75] making use DDPV and ADPV curves of electro-reductions with k0-values
of the nonlinear Matyushov solvation model.[76] within the range 10¢3–10¢4 cm s¢1 and very small transfer coef-
According to all of the above, for solution-phase redox cou- ficients (a < 0.3, see Figure 6) predicted at macroelectrodes[85]
ples, the simpler, 3-parameter BV model can be recommended have also been found at microelectrodes,[77c] though it gradual-
for the fitting of experimental data, complemented with the ly disappears as the electrode size is reduced. Note that in the
physical insights derived from the asymmetric Marcus model. case of electro-oxidation processes the splitting is predicted
On the other hand, the analysis of surface-bound redox cou- for large a-values (a > 0.7). It is also worth mentioning that the
ples should be performed using the asymmetric Marcus–Hush splitting is not predicted by the symmetric Marcus–Hush kinet-
model, which is the only theoretical approach (among those ic model.[70a]
considered here) compatible with all the experimental results In practice, deviations from the fully-reversible behavior can
reported in the literature. be detected by comparison of the experimental results with

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 233 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
pulse amplitude in DDPV are the same for reversible systems,
as well as the heights of the maximum (IM) and minimum (IM)
in the ADPV signal ([Eqs. (20), (21)] and Figure 6).[88] These be-
haviours do not hold for finite-kinetic electrode reactions.[88]
Note that this reversibility criterion may not be conclusive for
systems where the diffusivities of the oxidised and reduced
species differ significantly. In such cases, even if the electron
transfer is reversible, the position of the DDPV and ADPV sig-
nals depends on the double pulse duration and the electrode
size and the values of j DIpeak(DE < 0)/DIpeak(DE > 0) j and j IM/Im j
differ from 1.
The quantification of the electrode kinetics is possible from
single-point fitting of the DDPV and ADPV curves.[77a–c] For this,
the peak height (in DDPV) and maximum current (in ADPV) are
more sensitive in the case of quasi-reversible processes where-
as the DDPV peak potential and the ADPV crossing potential
are more appropriate for irreversible electrode reactions. Thus,
from the fitting of the variation of the DDPV peak current and
potential or the ADPV maximum current and crossing potential
with the duration of the double pulse (t1 + t2), the kinetics of
three electro-reduction processes of different reversibility were
successfully determined using mercury micro-hemispheres as
working electrodes: 3-nitrophenolate anion in DMSO, 3-nitro-
phthalate di-anion in DMSO, and europium(III) in H2O.[77b] The
analysis of the “first pulse” and “second pulse” components of
the DDPV curve has also been recently proposed for the inves-
tigation of the electrode kinetics.[89]
It is worth noting that the influence of the kinetic parame-
ters on the response in differential multi pulse voltammetry
(DMPV) is qualitatively analogous to that discussed above for
DDPV, and that the fitting methodology proposed is also appli-
cable. Nevertheless, quantitative kinetic analysis of the DMPV
Figure 5. Illustration of the effects of the thermodynamics and kinetics of curves requires for the use of numerical simulation methods.[90]
coupled homogeneous chemical reactions on differential double pulse vol-
Thus, although the DMPV method is generally preferred to
tammetry (DDPV)[79a] (A) and reverse pulse voltammetry (RPV)[80] (B). Grey
solid lines correspond to a simple reversible E mechanism. DDPV given that equilibrium conditions are not recovered
after each pair of pulses and so the time of experiments is
shorter, the theoretical modeling and analysis of results are
those predicted for fast electron transfers with equal diffusion more complex due to accumulative effects. Only for reversible
coefficients for the reduced and oxidised species: electrode processes or at ultramicroelectrodes are analytical
DDPV solutions available for DMPV.[91] In the case of nonreversible
processes at planar electrodes or conventional microelectro-
Epeak ¼ Ef0 des, the superposition principle is not applicable due to the
” ”  Ž ð20Þ
”DIpeak ” ¼ FADcŸ fG ðt2 Þtanh F jDEj time-dependence of the surface concentrations, and numerical
A RT 4
methods must be employed to simulate and fit the DMPV sig-
nal.[3a]
ADPV

3.1.2 Reverse pulse voltammetry (RPV)[92]


Ecross ¼ Ef0
” ” ð21Þ The RPV response under transient conditions shows a cathodic
” IM ”
”Im ” ¼ 1
and an anodic branch (without requiring the initial presence of
the product species), and the shape of the RPV curve is greatly
with fG(t2) being a time function, the form of which depends affected by the electron transfer kinetics as shown in Figure 7.
on the electrode geometry.[86] Note that in Equation 20, it is As k0 decreases, the RPV voltammogram gradually splits into
considered that the arithmetic average (E1 + E2)/2 is chosen for a cathodic and an anodic wave. Also, when the process is slug-
the x-axis potential.[87] gish (k0 < 10¢3 cm s¢1) and the second potential pulse is long
With respect to the peak currents, the magnitude of the enough (t2 t1), a maximum (“bump”) is observed in the
peaks obtained with positive (DE > 0) and negative (DE < 0) anodic wave that is more apparent at large electrodes.[77d] With

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 234 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 6. Influence of the Butler–Volmer kinetic parameters on the response of a one-electron reduction reaction in DDPV (left) and ADPV (right) at a (hemi)-
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffi
spherical microelectrode (r0 = 30 mm). t1 = 1 s, t1/t2 = 50, DE = 50 mV. E1,2 = (E1 + E2)/2, Id,p(t2) = FAcA* D= pt 2 .

regard to the transfer coefficient, the a-value affects both the netics, and so they allow for simultaneous determination of
position and slope of the cathodic and anodic branches such the diffusion coefficients of both electroactive species. For this,
that the cathodic wave is steeper and shifts to smaller overpo- an electrode of appropriate size must be employed (in the
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffi
tentials as a takes larger values (Figure 7), the opposite being case of spherical electrodes: 2 Dt1 > r0 > 0:8 Dt1 ), not being
true for the anodic wave. Therefore, visual inspection of the possible either at macro- or at ultramicro-electrodes unless the
RPV curve enables us to estimate the electrochemical reversi- two electroactive species are initially present.[93] Therefore, the
bility of the system, as well as the transfer coefficient. A sum- use of microelectrodes of medium size in combination with
mary of the reversibility criteria for DDPV, ADPV, and RPV is the RPV technique enables the determination of the diffusion
found in Table 1. coefficients and the study of the electrode kinetics in a single
Analogously to the cases of DDPV and ADPV, it is possible to experiment.
quantify the electrode kinetics parameters in RPV from the fit-
ting of “singular” points of the curve. Thus, the values of the
3.2 Square wave voltammetry (SWV)
mid-wave potentials of the cathodic and anodic branches to-
gether with their variation with the double pulse duration Square wave voltammetry is well-known for its high sensitivity
have been employed with success for the kinetic study of the in electroanalysis, and it is also a powerful technique in the
electroreduction processes mentioned in Section 3.1.1. study of electrode kinetics and reaction mechanisms of solu-
It is also worth noting that the cathodic and anodic limiting tion-phase and surface-confined redox systems.[94] SWV in-
currents of the RPV curves are not affected by the electrode ki- cludes the benefits of differential pulse techniques along

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 235 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 7. Influence of the Butler–Volmer kinetic parameters on the response of a one-electron reduction process in RPV at a (hemi)spherical microelectrode
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffi
(r0 = 30 mm). t1 = 1 s, t1/t2 = 10. Id,p(t2) = FA cŸA D= pt 2 .

Table 1. Reversibility criteria for the DDPV,[88] ADPV, and RPV techniques for electro-reduction processes when the diffusion coefficientspof ffiffiffiffi the electroactive
E þE pffiffiffiffiffiffiffi
species are equal. Note that in DDPV:Eindex ¼ 1 2 2 . The form of the functions fG(t) depends on the electrode geometry.[86] Id,p(t2) = F A CAŸ D= pt2 .

Fully reversible Quasi-reversible Fully irreversible


0 0
DDPV * Epeak ¼ E f
* Epeak < E f
* Epeak < Ef0
” ” ” ” ” ”
”DIpeak ðDE<0ޔ ”DIpeak ðDE<0ޔ ”DIpeak ðDE<0ޔ
* ”DI ðDE>0ޔ
peak
¼1 * ”DI ðDE>0ޔ6¼1
peak
>1
* ”DI ðDE>0ޔ
peak
DIpeak DIpeak
 Ž * I ðt Þ ¼ gðt1 þ t2 Þ * I ðt Þ 6¼gðt1
þ t2 Þ
” ” d;p 2 d;p 2
”DIpeak ” ¼ FADcŸ fG ðt2 Þtanh F jDEj at macroelectrodes for a given t1/t2 value at macroelectrodes for a given t1/t2 value
* A RT 4

ADPV * Ecross ¼ Ef0 * Ecross < Ef0 * Ecross < Ef0


” ” ” ” ” ”
”IM ” ”IM ” ”IM ”
* ”I ”
m
¼1 * ”I ”6¼1
m
* ”I ”
m
>1
” ” ” ” ” ”
”IM ” ”IM ” ”IM ”
* ”I ”6¼f ðt; r0 Þ * ”I ” ¼ gðt1 þ t2 Þ * ”I ”6¼gðt1 þ t2 Þ
m m m
at macroelectrodes for a given t1/t2 value at macroelectrodes for a given t1/t2 value

RPV * One wave * Transition between one and two waves * Two waves
* Bump when t2 t1

with those of potential sweep methods (fast experiments). Y peak;disc ¼ 0:5141 þ 0:9047x
The effects of the BV kinetic parameters on the SWV peaks Y peak;sph ¼ 0:5183 þ 0:7501x
are similar[94a, 95] to those described in Section 3.1.1 for DDPV ð22Þ
Y peak;cyl ¼ 0:5207 þ 0:2953x ¢ 0:008z2
such that the peaks are smaller, broader, and situated at
larger overpotentials as k0 decreases and a decreases (in the Y peak;band ¼ 0:5189 þ 0:3602x ¢ 0:007z2
case of electro-reductions) or a increases (for electro-oxida-
tions). pffi pffiffiffiffi
I t Dt
Simple and rapid diagnosis criteria for the detection of finite where Y ¼ FA SWpffiffiffiffiffi
DcŸ
and z ¼ qG , with t being half the square
G 0

electrode kinetics can be established based on deviations from


the SWV signal expected for fully-reversible processes. Thus, wave period (t = 1/2 f) and qG the characteristic dimension of
the value of the peak current of nonreversible processes will the electrode: the radius for discs, spheres and cylinders and
be smaller than that predicted by the following expressions for the half width for bands.
reversible electrode reactions at disc, (hemi)spherical, band and The value of the half-peak width (W1/2) is also a parameter of
cylindrical electrodes and microelectrodes under typical SWV interest given that the W1/2-value for one-electron reversible
conditions (ESW = 50 mV, Es = 5 mV):[86b] processes is defined only by the SW amplitude (ESW, [Eq. (23)],

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 236 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
W1/2 being independent of the electrode geometry and fre- satisfactory results to the study of solution-phase and surface-
quency employed:[86b] confined[3a, 98]redox systems.
Cyclic and reverse scan SWV has also been proposed and
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi! employed in quantitative kinetic studies.[69, 71, 99] In the case of
RT 1 þ e2hSW þ 4ehSW þ ð1 þ e2hSW þ 4ehSW Þ2 ¢4e2hSW quasi-reversible and irreversible diffusional processes
W1=2 ¼ ln pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
F 1 þ e2hSW þ 4ehSW ¢ ð1 þ e2hSW þ 4ehSW Þ2 ¢4e2hSW (Figure 9), a double peak is observed in the reverse scan at
negative (cathodic nature) and positive (anodic nature) poten-
ð23Þ
tials with respect to the formal potential. The splitting of the
peak is more apparent as the electrode size and/or the fre-
FE
where hSW ¼ RTSW is the dimensionless SW amplitude. Thus, in quency increase and they separate as the electron transfer is
absence of ohmic drop effects,[96] experimental W1/2-values more sluggish (Figure 9). The fitting of the peak potentials,
larger than those predicted by [Eq. (23)] indicate a nonreversi- peak widths, and relative peak height enables the characterisa-
ble behaviour. The effect of the step potential (Es) on the SWV tion of the electrode kinetics. This approach has been applied
signal also offers a simple criterion to estimate the degree of to the study of the electroreduction of 2-nitropropane[71] and
reversibility.[97] Thus, whereas the SWV response of reversible europium (III)[99a] on mercury electrodes and microelectrodes.
systems is scarcely affected by Es, the SWV response of irrever-
sible reactions varies significantly with Es : the smaller the Es
value, the smaller the SWV peak. 3.3 Cyclic pulse voltammetries
For quantitative analysis of the SWV response of nonreversi- Software packages of modern electrochemical instrumentation
ble electrode processes, numerical simulation methods are enable the researcher to “customize” the voltammetric pertur-
necessary,[3a] semi-analytical solutions in the form of a system bation applied to the system. Within this context, Jadresko
of recursive formulae being also available.[94a] Though frequen- et al.[100] have recently proposed two new variants of pulse
cy-based approaches have been usually considered for the in- techniques: cyclic multi pulse voltammetry (CMPV) and cyclic
vestigation of the electrode kinetics,[94a] Mircevski et al.[94i] have differential multi pulse voltammetry (CDMPV). The potential-
recently proposed a new approach based on the variation of time program is analogous to that employed in normal/reverse
the SW amplitude rather than the time scale of the scans. Ac- pulse voltammetry and DDPV, respectively, with the key differ-
cording to the new amplitude-based strategy, the electrode ki- ences that the perturbation is applied in a “cyclic mode”, and
netics is characterised from the variation of the separation of that equilibrium conditions are only restored at the end of the
the peak potentials of the forward and backward components experiment. As a result, the characterisation of the system is
of the potential-corrected SW voltammogram and/or the peak more complete and sound, and the electrochemical measure-
current of the net response with the SW amplitude (ESW). This ments are faster. The resulting signal enables qualitative analy-
variation is sensitive to the kinetic parameters as shown in sis of the process (including the electrode kinetics) from visual
Figure 8 for the amplitude-normalised peak current (DIpeak/ESW). inspection of the voltammograms as well as quantitative analy-
The amplitude-based methodologies have been applied with sis from single-point fittings.[100]

Figure 8. Influence of the electrode kinetic parameters on the amplitude-based


pffiffiffiffiffi quasi-reversible maximum of a one-electron reduction at a hemispherical mi-
croelectrode corresponding to a solution-phase redox system. w ¼ k0 = fD .

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 237 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 9. Influence of the electrode kinetic
pffiffiffiffi
ffi parameterspffiffiffion
ffi pthe
ffiffiffiffiffi reverse scan of the cyclic SWV response of a one-electron reduction corresponding to a solu-
tion-phase redox system. w ¼ k 0 = fD , Id;p ðtÞ ¼ FAcŸA D= pt .

3.4 Staircase versus analogue cyclic voltammetry

Since the early 90s most commercial potentiostats have been


predominantly digital, computer controlled devices. In part
due to cost and ease of implementation the basic cyclic vol-
tammetric technique provided by these devices involves the
application of a ‘staircase’ ramping potential. The use of a stair-
case waveform for voltammetry was initially proposed as
a route by which Faradaic and capacitive currents may be
more readily experimentally discriminated between (cf. voltam-
metric pulse techniques).[101] This discrimination is partially ena-
bled on the basis of the differing time constants associated
with diffusional redox (t¢0.5, for a linear mass-transport regime)
and capacitive charging (e¢t, in the heavily simplified RC circuit
analogy) processes. However, prima facie there is no reason to
assume that staircase and analogue cyclic voltammetry are
equivalent.
Figure 10 depicts the variation of the potential used for Figure 10. Comparison of the voltage wave forms used for staircase (red)
‘staircase’ (red) and true analogue (black) cyclic voltammetry. and true analogue (black) cyclic voltammetry. Zoomed inlay depicts an indi-
vidual step showing the sampling alpha scale. When a = 1 the current is
For a given electrochemical system studied via analogue cyclic
sampled at the end of the step; alternatively, a = 0 implies a current mea-
voltammetry (CV), the measured response is simply a function surement at the beginning of the step. Data depicts the wave form used for
of the scan rate (assuming appropriately chosen start, finish, a cyclic voltammogram (0–1 V) at a scan rate of 0.1 V s¢1 and with a step po-
and turning potentials). Conversely, for staircase cyclic voltam- tential of 20 mV.
metry (SCV), the resulting voltammogram is a function of the
scan rate (step height/step time = V s¢1), the step size (Estep/V), diffusional redox species the use of SCV tends to lead to vol-
and the point (or points) at which the current is sampled tammetric waves that exhibit larger peak-to-peak separations
during each step. In the case that the current is sampled once and suppressed peak heights. It should be recognised that all
during each step the time at which the current is sampled is analytical expressions (the Randles–Ševč†k equation for exam-
expressed as the dimensionless value alpha (a), where an ple) and commercially available simulation packages assume
alpha value of one or zero implies the current is sampled at the utilisation of an analogue potential ramp. Consequently,
the end or beginning of each step respectively (see inlay of the use of these equations or simulation software for quantifi-
Figure 10). This sampling alpha value bears no relation to the cation of SCV results can lead to erroneous results.
transfer coefficient and the two should in no way be conflat- For a reversible diffusional process, equivalency (within ex-
ed! perimental error; peak current, Ip error < 3 %, peak position
In the late 80s, Osteryoung published a series of papers in- within 2 mV) between SCV using an alpha value of 1 and ana-
vestigating the differences and possible equivalences between logue CV reportedly requires the use of a step size of
staircase and analogue cyclic voltammetry.[102] Importantly, for 0.26 mV.[102a] However, for diffusional processes this constraint

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 238 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
may be relaxed through the use of an alpha value of 0.25–
0.3,[103] enabling the use of slightly larger step potentials with-
out too significant a deviation from the results predicted for
analogue CV. Depending on the experimental conditions and
the equipment used, it may or may not be possible to select
experimental parameters that allow the SCV response to close-
ly approximate that obtained from analog CV.3 Although the
response is improved by using an alpha of 0.3, for situations in
which accuracy is highly pertinent it may be advisable to
revert to using true analogue cyclic voltammetry;[104] alterna-
tively one may explicitly simulate the response accounting for
the staircase ramping potential.[105] Recent theoretical studies
have investigated the influence of the alpha value in staircase
voltammetry for the case in which the diffusion profile at an
electrode is transitional between the linear and convergent
limits.[105] Moreover, expressions for the analytical solution at
a microelectrode for the staircase response of single-, multi-,
and catalytic electron transfer processes have also been provi-
ded.[81, 86a, 106]
The above discussion has focused on the voltammetric re-
sponse of diffusional redox processes, where for many systems
the application of SCV yields qualitative but not quantitative Figure 11. Staircase voltammograms of yeast cytochrome c peroxidase (CcP)
correspondence with the analogue technique. In contrast, for at different sampling times. Conditions: 0.13 mm CcP in 20 mm sodium ace-
tate, pH 5.45, 0 8C, n = 20 mV s¢1, Estep = 5.04 mV, and a = 0.13, 0.15, 0.18,
surface-bound species, the obtained voltammetric results can
0.25, 0.35, 0.48, and 0.90. Each trace is the average of four cycles. Adapted
differ profoundly between the techniques. In the most extreme with permission from Ref. [107]. Copyright 1999, American Chemical Society.
cases, where the surface species exhibits fast electron transfer
kinetics, it may arise that the Faradaic charge transfer occurs
prior to the measurement point on the step. In this situation
the use of SCV may yield a voltammogram that is completely ously over the course of a potential step and averaging the
devoid of a voltammetric feature even if the redox species is result; for surface bound species this results in a voltammo-
present. To exemplify this point, Figure 11 shows the staircase gram closely comparable to that found with the use of ana-
voltammetric response of cytochrome c peroxidase on a pyro- logue CV. Depending on the potentiostat manufacturer, this
lytic graphite electrode, where the experimental sampling technique goes by a variety of names including ‘current inte-
alpha value has been varied between 0.13–0.9.[107] gration’ and ‘surface mode sampling’. Finally, when investigat-
For situations in which the current is sampled towards the ing the fundamentals of the electron transfer process of sur-
end of a step, the reversible cyctochrome c peroxidase voltam- face-bound species by voltammetric techniques, if a staircase
metric wave is not recorded. SCV of a surface-bound feature potential ramp has been used, it is imperative that this is
only becomes equivalent to analogue CV when the scan rate taken into account in the simulations so as to ensure validity
(v) is greater than 10 k0Estep.[107] Subsequently, the use of stair- of the results.[110]
case voltammetry for the investigation and quantification (in
terms of surface coverage) of a reversible surface bound pro-
cess must be approached with caution. To this end it is noted
3.5 Insights into the concept of the diffusion layer thickness
that one of the prime examples of such a system ‘misrepre-
sented’ by SCV is encountered with hydrogen UPD on plati- The Nernst diffusion layer concept (or linear diffusion layer)
num,[108] however other molecular species can encounter simi- provides a useful approach to the species concentration pro-
lar problems.[109] This is particularly true when investigating sur- files and the diffusive mass transport in electrochemical sys-
face-bound species as a function of temperature, where it may tems. The thickness of such a layer, d, informs about the
be found that at higher temperatures the surface-bound extent of the region in solution where concentration changes
redox-wave is essentially ‘lost’ when using SCV. This can occur take place and the efficiency of diffusion under given experi-
simply due to the increase in the electron transfer kinetics as mental conditions. This information is essential in digital simu-
a function of temperature. One method by which these prob- lation of electrochemical experiments,[3a] the evaluation of pos-
lems may be circumvented is by sampling the current continu- sible interferences due to convective mass transport[111] and
double layer effects,[112] and the design of micro- and nanoelec-
3
trode arrays in order to predict the overlapping between adja-
It is specifically noted that the ability to alter the experimental alpha value
cent diffusion domains.[113] Nevertheless, only very recently
was possible with Autolab’s earlier software package GPES (Metrohm,
Utrecht, Netherlands); however, this functionality was removed with the intro- have the effects of finite electrode kinetics and convergent dif-
duction of NOVA where a default value of 1 is used. fusion on d been investigated.[114] Thus, analytical expressions

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 239 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
have been reported for the study of the linear diffusion layer From Equations 24–26, the following expressions are ob-
thickness in any voltammetric experiment. tained for the linear diffusion layer thickness at planar, (hemi)-
spherical, and cylindrical electrodes in any voltammetric tech-
nique consisting of a sequence of p potential pulses of the
same duration, t:
3.5.1 Nonplanar diffusion in any voltammetric technique
In the case of uniformly accessible electrodes, such as (hemi)- 1
spheres and cylinders, the diffusion problem can be reduced dpplanar ¼ ð1þeh Þ P p Zm ð27Þ
pffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi

p

m¼1 p¢mþ1
to a single spatial coordinate corresponding to the direction pDt

normal to the electrode surface, q (Figure 12). The linear diffu- 1


sion layer thickness is defined as the distance to the electrode dpsph ¼ 1 ð1þe Þ hp Pp Z ð28Þ
þ pffiffiffiffiffi
ffi m¼1
ffiffiffiffiffiffiffiffiffiffiffi
p m
surface where the linear concentration profile takes the bulk r0 pDt p¢mþ1

value c* (Figure 12 a). Accordingly, the d value is given by:


1
dpcyl ¼ Pp ¨ ¦ ð29Þ
Ÿ ð1 þ ehp Þ Z f
m¼1 m cyl tmp
c ¢ cA;surf
d¼ Ž A
ð24Þ
@cA ðq;t Þ
@q qsurf where:
( 1
Taking into account that for a fully reversible electron trans- 1þeh1 for m ¼ 1
Zm ¼ ð30Þ
fer A + e¢Ð B, with species A and B having equal diffusion co- 1
¢ 1þeh
1
for m > 1
1þehm m¢1

efficients, the surface concentrations only depends on the ap-


plied potential, Ep, such that when cB* = 0 it is fulfilled that: At microdiscs and microbands, the mass transport of species
in solution towards/from the electrode is not uniform over the
cŸA whole electrode area. Thus, the mass flux is higher at the elec-
cŸA ¢ cA;surf ¼ ð25Þ
1 þ ehp trode edge than at the electrode centre, and the linear diffu-
sion layer thickness has an average character. After the applica-
tion of a sequence of p potential pulses of duration t, the aver-
F ðEp ¢E 0 Þ
where hp ¼ RT f , and that the surface concentration gradi- age linear diffusion layer thickness for a reversible process is
ent can be expressed as follows after a sequence of p potential given by:
pulses:
‡p ¼ 1
ddisc Pp ¨ ¦ ð31Þ
 Œ p h i ¨ ¦ ð1 þ e Þ hp
Z f
m¼1 m disc tmp
p
@c ðq; t Þ X
m¢1
A
¼ cA;surf ¢ cm
A;surf f tmp ð26Þ
@q 1
qsurf m¼1 ‡p ¼
d Pp ¨ ¦ ð32Þ
band
ð1 þ ehp Þ Z f
m¼1 m band tmp
where c0A;surf ¼ cŸA , tmp = (p¢m + 1)t and f(tmp) is a time function,
the form of which depends on the shape of the electrode em- Equations 27–29 and 31–32 enable the study of the behav-
ployed.[86] iour of the linear diffusion layer for very different electrodes

Figure 12. A) Schematic of the linear and real diffusion layer thicknesses in uniformly accessible electrodes. B) Dimensionless linear diffusion layer thick-
ness (red line) and dimensionless average linear diffusion layer thickness (black line) at disc electrodes.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 240 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
!
geometries and voltammetric techniques. Figure 13 shows the 2
1 ¢ cpffiffip FðcÞ
irrev
evolution of the (average) linear diffusion layer thickness in dsph ¼ r0 r0 kred ð1þeh Þ 2
ð33Þ
1þ D
ffiffi FðcÞ
p
chronoamperommetric (Figure 13 a) and linear sweep voltam- c p

metry (Figure 13 b) experiments at microelectrodes of different


shape (Figure 13 a ) and different radii (Figure 13 b). In both ex- where:
periments, the thickness of the linear diffusion layer increases
as the experiment proceeds and so the duration of the pertur- pffiffiffi
p  c Ž2 cŽ
bation. Regarding the electrode shape (Figure 13 a), for a given FðcÞ ¼ c exp erfc ð34Þ
r0, the d values coincide for any geometry at very short times 2 2 2
when diffusion is predominantly planar, with differences be-
pffiffiffiffiffi Š ‰
tween them becoming more apparent with time. Thus, d de- 2 Dt rk
c¼ 1 þ ð1 þ eh Þ 0 red ð35Þ
creases in the order: cylindrical > band > spherical > disc, which r0 D
means that the mass-transport efficiency (current density) fol-
lows the inverse order. With respect to the influence of the
kred ¼ k0 e¢ah ð36Þ
electrode size (Figure 13 b), the thickness of the linear diffusion
layer in absolute terms decreases as the electrode shrinks,
though the thickness relative to the electrode radius (i.e. d/r0) The degree of reversibility has a profound influence on the
increases, and it tends to 1 at microelectrodes. species surface concentrations, which are time-dependent for
It is also important to mention that d must be taken cau- nonreversible processes (unlike for fast electron transfers,
tiously as an estimation of the real diffusion layer thickness [Eq. (25)]):
given that these two magnitudes diverge very significantly at
microelectrodes and nanoelectrodes. Thus, whereas the ratio
 Œ
dreal/d is found to be about 2 at macroelectrodes (planar diffu- r0 kred cŸA 2
cŸA ¢ cA;surf ¼ h i 1 ¢ pffiffiffi FðcÞ ð37Þ
sion), it is about 15 at conventional microelectrodes (a few mi- r k
D 1 þ ð1 þ eh Þ 0 Dred c p
crometer-radius), and it tends to 100 at ultramicroelectrodes
(steady-state conditions).[114b]
as well as on the behaviour of the linear diffusion layer thick-
ness, which is potential-dependent as can be inferred from
Equations 33–36.
The results obtained from Equation 33 shows that the linear-
diffusion layer thickness of nonreversible processes is smaller
3.5.2 Finite electron transfer kinetics 1 1
than for reversible electron transfers (drev pffiffiffiffiffi
sph ¼ pDt þ r0 ), except
For the evaluation of the impact of the electrode kinetics on under plimiting
ffiffiffiffiffiffiffiffi current conditions and at ultramicroelectrodes
the linear diffusion layer thickness, the following expression (r0 ‚ pDt ) where dsph = r0 for any electrode kinetics, applied
has been deduced for processes of any degree of reversibility potential, and electrochemical method. In any other situation,
in single potential-step chronoamperometry at (hemi)spherical dirrev
sph is smaller than drev sph and it varies in the range:
2 rev
electrodes of any size:[114b] rev
p dsph < dsph < dsph .

Figure 13. Evolution of the (average) linear diffusion layer thickness in A) single-step chronoamperometry at different microelectrode shapes and B) linear
sweep voltammetry at spherical electrodes of different radii at 100 mV s¢1.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 241 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Section 4: Diffusion and Random Walks 1 X ¨ ¦
cð~
r; t ¼ 0Þ ¼ d~
r ¢~
p0;i ð39Þ
NA i

4.1 Modelling diffusional processes


where NA represents the Avogadro constant and dðÞ the Dirac
When it comes to the experimental validation of reaction delta function. For t > 0 the Dirac delta function, which de-
mechanisms (such as the mechanisms discussed in Sections 1 scribes the exact initial positions, can be replaced by the
and 2), digital simulations are a frequently used tool in today’s Green’s function G of the linear differential operator dt þ aD
electrochemical and electroanalytical research. Since simula- with a 2 R, which corresponds to the differential operator in
tions can be specifically designed to predict experimental data the diffusion equation. This Green’s function is given by:
of a certain electrochemical system based on a number of dif-
ferent models of underlying fundamental processes, they may  Œ
1 r2
provide data for direct comparison with experimentally-ob- Gðt Þ ¼ qðt Þ exp ¢ ð40Þ
ð4pat Þ3=2 4at
tained results. Herein, employed simulations always combine
two models: a model for charge-transfer processes at solid–
where qðt Þ is the Heaviside step function. We then obtain
liquid boundaries and a model for the mass transport of the
a continuous concentration profile as a function of time:
analyte. Such interface processes may, for instance, include
electrochemical interactions according to kinetic models such
” ”2 !
as the above-discussed Butler–Volmer or Marcus–Hush models, 1 X 1 ”~ p0;i ”
r ¢~
cð~
r; t ¼ 0Þ ¼ exp ¢ ð41Þ
or other physiochemical processes like adsorption and desorp- NA i ð4pDat Þ3=2 4Dt
tion kinetics, which provide the boundary conditions for the
mass-transport problem. In the common case that convective which again provides a probability density of the particles’ po-
processes are negligible, mass transport can be modelled sitions instead of desired discrete positions. However, in con-
through Fick’s second law:[115] trast to the direct solution of Fick’s second law, this result de-
scribes the temporal evolution of the concentration profile
with defined initial positions for each modelled molecule.

@cð~
r; t Þ In order to transform this finding into an exact distribution
¼ DDcð~
r; t Þ ð38Þ
@t of molecules, the average displacement jd~ r j of a particle after
a given time dt is calculated from the Green’s function. Via the
investigation of the mean squared displacement of an individ-
ual particle, we obtain:
where c is the concentration and D the isotropic diffusion coef-
ficient of the analyte. This equation can be solved via pffiffiffiffiffiffiffiffiffiffi
a number of different methods, most prominently through jd~
rj ¼ 6Ddt ð42Þ
finite differences[3a] or finite elements,[3c] which both provide
solutions for the concentration profile cð~ r; t Þ. A so-obtained so- for the three-dimensional case. In the one-dimensional case,
lution for the concentration profiles of reacting species then we calculate:
allows the calculation of the expected average current across
pffiffiffiffiffiffiffiffiffiffi
all electrochemical interfaces from the concentration gradient dxi ¼ 2Ddt ð43Þ
at the respective interface, which is the desired result in most
applications. However, the system’s intrinsic noise characteris- from the mean squared displacement.[116]
tics cannot be directly modelled through finite-difference or The temporal evolution of exact particle positions that fulfil
finite-element approaches as such noise characteristics are due the diffusion equation can hence be found by substituting
to the discrete nature of the analyte, which results in a stochas- each particles spatial probability density function by random
tic charge transfer across the interface, being particularly rele- displacement after discrete time steps of the width dt. Mathe-
vant at low concentrations or small structure sizes. matically, this approach can be expressed as:
Concentration profiles can rather be interpreted as probabili-
ty densities of finding a particle at a certain position, but do !
not allow direct insights into the stochastic nature of the 1 X Xn
cð~
r; t ¼ ndt Þ ¼ d ~
r ¢~
p0;i ¢ eij jd~
~ rj ð44Þ
charge-transfer process at the interface. One way to overcome NA i j¼0

this issue is the use of the random walk method. In this ap-
proach, the pathways, as well as all electrochemical interac- where ~ eij is a random unit vector. For reasons of computational
tions, of each analyte molecule are modelled individually. Ini- simplicity, however, the distribution of particles in two- or
tially at t = 0, all analyte molecule i positions ~ p0;i are set by ran- three-dimensional systems are often expressed in terms of in-
domly distributing molecules according to the initial concen- dependent one-dimensional average displacements, dxi , in
tration profile within the simulated space. The concentration each dimension. Based on this assumption, the previous equa-
profile can then be written in the form: tion is transformed to:

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 242 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
!
1 X Xn X 3 surface and the movement back to the initial position of this
cð~
r; t ¼ ndt Þ ¼ d ~
r ¢~
p0;i ¢ eijk dxk ð45Þ step. Mathematically, however, the expected time of such a re-
NA i j¼0 k¼1
flection, dt0, does not equal the time, dt, as it is presumed in
the formulation of the boundary condition. Using Equation 43,
where eijk is a one-dimensional random unit vector featuring we obtain:
the values + 1 or ¢1. Since this result provides exact stochastic  Œ
positions of all active molecules at any time, the noise charac- 1 dxi 2 dxi2
dt 0 ¼ 2 6¼ ¼ dt ð46Þ
teristics of the modelled system can be simulated in great 2D 2 2D
detail and in addition to the expected average values that can
be obtained from finite differences or finite elements. Such a boundary condition hence induces an error that
While random walk simulations offer the advantage of the scales with the spatiotemporal step width of the random walk.
ability of noise modelling, which is not offered by many other In order to reduce this error, the distance of boundary to the
methods, the random walk approach features two main disad- closest grid point of the random walk must then be corrected
vantages. First, since every molecules pathway has to be mod- to dx’ as it can be seen in Figure 14 B):
elled individually, the computational effort scales with the
number of active molecules in the modelled system. The 0 dx
random walk approach is hence not suitable to model high dxi ¼ pffiffiiffi ð47Þ
2
concentrations or large systems. Secondly, the appropriate def-
inition of boundary conditions may be difficult as the above as was discussed by K•telhçn et al.[117] If, however, computa-
discussed theoretical justification of the random walk approach tional effort is not a limiting factor of the simulation, the sim-
implies a significant limitation: The Green’s function approach plest way to circumvent the problem of the definition of ap-
chosen in Equation 41 solely describes the temporal evolution propriate boundary conditions to choose a sufficiently small
of a diffusing particle in the absence of diffusion boundaries. spatial step width for the random walker. Since the induced
In order to model a real electrochemical set-up including elec- error scales with the spatiotemporal step width of the simula-
trodes and inactive surfaces, approximations must be made. tion, the deviation from the analytical result will scale with dt
To illustrate this problem, we focus on the common case of and dxi . When simulating a large number of molecules or
a random walker on a one-dimensional grid in between two a long experiment, this leads to a significant increase in com-
reflecting boundaries; the approach is, however, equally appli- putational effort and is therefore often not applicable.
cable to three dimensions. In the one-dimensional case, the In recent years, the noise modelling capabilities of the
boundary condition at a boundary can be formulated in multi- random walk approach have been exploited in a number of
ple ways; the most common definition can be seen in Fig- different studies focusing on electrochemical systems. These
ure 14 A). The closest grid point to the boundary is separated studies include more fundamental analyses of stochastic
versus statistic descriptions of diffusion processes[118] as well as
the description of experimental systems. Cutress et al. for in-
stance used a GPU-based random walk simulation to investi-
gate cyclic voltammetry[119] and potential-step chronoamper-
ometry[120] at low concentrations, while K•telhçn and Compton
focused on the noise-characteristics of the mediated Faradaic
current across a nanoparticle impacting on a Faradaically inac-
tive electrode surface.[121] Aside from the modelling of systems
at low concentrations, random walk simulations can further be
employed to investigate noise characteristics of electrochemi-
cal sensors in nanofluidic devices. Hereby, applications include
the modelling of methods for single molecule detection[122] as
well as the simulation[117, 123] of spectra obtained from electro-
chemical correlation spectroscopy (ECS).[124]

Figure 14. Examples of two different definitions of boundary conditions in


4.2 From molecules to nanoparticles
a random walk simulation. The grid in (B) is displaced relative to the boun-
Aside from its application in modelling diffusion of molecular
dary.
probes, the random walk approach can be used to simulate
the Brownian movement of nanoparticles.[125] Here, random
from it by dxi =2 and, during each temporal step dt, the walks are particularly helpful, since experiments focusing on
random walker may either remain on this position or perform the mass transport or electrochemistry of nanoparticles, such
a step away from the boundary. In this case, the reflection at as nano-impacts, are usually performed at concentrations that
the boundary has to be divided into two independent first are sufficiently low to resolve the reaction of individual nano-
passage problems: the diffusive movement to the boundary particles. Currents recorded in such experiments are hence

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 243 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
strongly influenced by the stochastic nature of the system, spend on average a significantly longer time in this zone than
which can be modelled through random walks. in any other zone of equal size in the bulk reservoir. This effect
When modelling the diffusion of nanoparticles, the above of hydrodynamic adsorption was recently discussed generally
discussed approach has to be slightly modified in order to ac- and with respect to the average time of residence that a cata-
count for the nanoparticles’ distinct diffusional characteristics lytically active particle spends within the zone of electron
that differ from molecular diffusion: Due to their greater size, transfer near an electrode.[125b]
nanoparticles are affected by the effect of near-wall hindered
diffusion when they approach a diffusional boundary. This
effect leads to an anisotropic diffusion coefficient distinguish-
Section 5: Modelling Migration and Diffusion
ing between diffusion perpendicular and in parallel to the The vast majority of electrochemical experiments are carried
boundary. The diffusion coefficient D? in the perpendicular out in the presence of a large excess of inert, fully dissociated
case is then given by: electrolyte.[78] The purpose of this electrolyte is to generate
a high ionic strength in solution, which will efficiently dissipate
D? 6h2 þ 2ah the excess charge necessarily introduced into solution via elec-
¼ 2 ð48Þ trolysis, and suppress the resultant electric field. There are two
D1 6h þ 9ah þ 2a2
main reasons that this is normally the case.
where a is the radius of the particle, h the particle’s elevation The first reason is to prevent ohmic drop.[26] The driving
from the surface, and D1 the bulk diffusion coefficient.[126] The force behind electron transfer is the potential difference be-
latter can be approximated well via the Stokes–Einstein equa- tween the electrode and the point in solution where electron
tion: transfer takes place, Ÿm þ Ÿs . The bulk solution, far from the
electrode, has some fixed potential Ÿbulk . If the potential drop
between Ÿm and Ÿbulk occurs over a distance greater than the
kB T
D1 ¼ ð49Þ electron tunneling distance for electron transfer (outside the
6pha
zone of electron transfer, ZET) then the full driving force will
where kB is the Boltzmann constant, T the temperature, and h not be felt; the potential difference is lowered as a result of
the viscosity of the solvent.[26] The parallel component of the ohmic drop. If a large amount of excess electrolyte is added to
diffusion coefficient D k can be described through:[127] efficiently dissipate excess charge, the distance over which the
drop between Ÿm and Ÿbulk occurs is compressed to a distance
much smaller than the ZET. This being the case, the electron
Djj 9  a Ž 1  a Ž3 45  a Ž4 1  a Ž5
¼1¢ þ ¢ ¢ transfer is then driven by the maximum potential difference,
D1 16 a þ h 8 aþh 256 a þ h 16 a þ h
Ÿm ¢ Ÿbulk . This is exemplified schematically in Figure 16.
ð50Þ The solid line shows the potential profile under conditions
of high support (a large excess of supporting electrolyte),
The distance-dependency of both diffusion coefficients can be which is compressed to short distances and does not extend
found in Figure 15. very far into the ZET. The dashed line shows low support con-
As it can be seen in the plot, perpendicular diffusion slows ditions (small amounts of supporting electrolyte), where the
down near the boundary and eventually vanishes at the sur-
face. Diffusing particles that are located in this area hence

Figure 16. Schematic representation of solution potential profiles under con-


Figure 15. Diffusion coefficients perpendicular and in parallel to a boundary ditions of high (solid line) and low (dashed line) support. The dotted lines
relative to the bulk diffusion coefficient, (D(h)_(? , j j)/D_1). Presented represent the zone of electron transfer extending from the electrode surface
graphs were for a particle featuring a radius of 15 nm.. out a certain distance into solution.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 244 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
electric field extends out beyond the ZET, resulting in a smaller 5.1.1 Theoretical treatments neglecting the double layer
driving force for electron transfer.
Secondly, a compressed electric field will eliminate migratory The electroneutrality approximation simply assumes the solu-
effects from the mass transport of solution phase species.[128] If tion to be electroneutral at all points:[130]
an electric field extends far beyond the electrode surface, the X
resulting potential gradient in solution will induce electrical mi- zi ci ¼ 0 ð53Þ
gration of charged species, in addition to diffusion. The majori- i

ty of analytical theory in electrochemistry assumes diffusion-


A potential profile which satisfies this condition is the mass
only conditions, and the presence of migration complicates
transport equation and the appropriate electron transfer boun-
matters. For these two reasons, excess supporting electrolyte
dary condition (for cyclic voltammetry this will be the Nernst
is usually added to an electrochemical experiment. However,
equation, the Butler–Volmer equation, or Marcus–Hush kinet-
migration effects can offer extra kinetic and mechanistic infor-
ics, as discussed above). This approximation greatly facilitates
mation unavailable at high support levels.[129] For this reason, it
analytical solution of the Nernst–Planck equation.[130] Where no
may be desirable to carry out electrochemical experiments
such analytical solutions exist and numerical simulation is re-
where a small amount, or zero, supporting electrolyte is added
quired, in the case of transient voltammetry for example, this
to solution. For such experimental cases new theoretical
approximation offers little advantage over other methods
models are needed to describe the experiments and allow the
which do not make the a priori assumption that the solution is
experimental electrochemist to interpret results.
electroneutral at all points.
A more rigorous method is the zero field approximation.[131]
By assuming that the electrical double layer is infinitesimally
5.1 The Nernst–Plank–Poisson equations small, the charge on the electrode surface is completely can-
If migration effects are present in an experiment, Fick’s second celled in a negligibly small layer of solution immediately adja-
law alone becomes inadequate to describe mass transport. In- cent to it. The electrode surface and this infinitesimal layer of
stead, the Nernst–Planck equation is used: solution taken as a whole then has zero charge, and hence
zero electric field exists at the electrode surface:
Š ‰
@ci zF @Ÿ
¼ Di r2 ci þ i ðrci rŸs þ r2 Ÿs Þ ð51Þ ¼0 ð54Þ
@t RT @x

where ci is the concentration of species i (mol m¢3), t is time This boundary condition may then be used with the Nernst–
(s), Di is the diffusion coefficient of species i (m2 s¢1), zi is the Planck–Poisson system of equations to numerically generate
charge on species i (multiples of the electronic charge), F is a potential profile and species concentration profiles across
the Farday constant, R is the gas constant, T is temperature (K) the solution. This method has been shown to be successful in
and Ÿs is solution potential (V). This equation describes the simulating diverse experimental data, at both micro and macro
mass transport of a solution-phase species in the presence of electrodes,[104, 129b–g, 132] and is used to obtain theoretical results
an electric field. discussed in sections 5.2 to 5.4.
The solution potential, fs, is described using the Poisson
Neither of these approaches models the electrical double
equation:
layer, and both assume it is small enough compared to the de-
pletion layer around the electrode to neglect. This approxima-
F X tion is generally valid for electrodes of radius greater than ap-
r 2 Ÿs ¼ ¢ zc ð52Þ proximately 10 mm. For electrodes smaller than this, the deple-
e0 es i i i
tion layer around the electrode will approach the size of the
electrical double layer, which can then no longer be neglected.
where e0 is the vacuum permittivity (F m¢1) and es is the rela-
tive permittivity of the solvent. These two equations together
5.1.2 Theoretical treatments including the double layer
constitute the Nernst–Planck–Poisson system of equations, and
subject to appropriate boundary conditions may be used to Various models exist to simulate voltammetry when a signifi-
model electrochemical experiments in the absence of excess cant double layer is present. The simplest assume that electron
supporting electrolyte. transfer takes place at a plane located at a fixed distance from
The Poisson equation can be applied to the simulation of the electrode surface, the plane of electron transfer (PET). This
weakly supported electrochemical experiments in different approach leads to the identification of the Levich and Frumkin
ways. The simplest models assume that the electrical double effects.
layer is negligible in extent beyond the electrode, and thus Levich predicted[133] that for a weakly supported system, if
completely exclude it. There are then two main approaches: zA n (where zA is the charge on the species undergoing electron
the electroneutrality approximation and the zero field approxi- transfer and n is the number of electrons transferred from the
mation. species to the electrode, which is positive for an oxidation and

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 245 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
negative for a reduction) is greater than zero, then the reacting in the simulations it is seen that even 100 times as much sup-
species is excluded from the electrode at large overpotentials porting electrolyte as electroactive species is not sufficient in
if a significant double layer is present, lowering the current. this case to exactly reproduce the fully supported result.
This leads to the prediction of peak-shaped steady-state vol-
tammetry. A lowering of the current is also predicted by the
5.3 The effects of weak support
Frumkin effect,[134] which predicts a reduced electrochemical
rate constant inside a double layer at large overpotentials, if Figure 17 also usefully demonstrates some key features of
zA n is greater than or equal to zero. weakly supported voltammetry. As the amount of supporting
The absence of experimentally observed peak-shaped electrolyte is lowered, the ohmic drop results in a larger peak-
steady-state voltammetry in many microelectrode systems sug- to-peak separation and a reduced current.
gests these pictures are not complete. Rather than assuming The effect of the charge born by the electroactive species
electron transfer to occur solely at the plane of electron trans- was investigated thoroughly by Belding and Compton.[137] Over
fer, Dickinson and Compton[135] developed a model where elec- the course of a reduction, negative charge is necessarily intro-
tron transfer occurs via a distance-dependent tunneling mech- duced into the solution, and without a large amount of sup-
anism across the diffuse double layer. This was shown to dra- porting electrolyte to dissipate this charge it builds up around
matically mitigate the Levich and Frumkin effects, and sigmoi- the electrode. Hence, positively charged species will be attract-
dal steady-state voltammetry is regained. ed towards the electrode, and negative species repelled away
from it. This is seen in Figure 18, where the positive species
has a higher limiting flux than the neutral species due to its
5.2 How much suppporting electrolyte is needed?
For steady-state voltammetry, it has been proposed that
a ratio of supporting electrolyte to electroactive species of 26
is sufficient to ensure full support.[136] Dickinson et al.[104] dem-
onstrated that for macroelectrode systems, this is not enough
supporting electrolyte and significantly more is needed to
avoid the effects of ohmic drop and migration becoming ap-
parent. A ratio in excess of 100 is shown to be required in
some cases. As well as large electrodes, it was shown that fast
scan rates and slow diffusion of the electroactive species (all of
which lead to more transient voltammetry), as well as slow dif-
fusion of the supporting electrolyte all necessitate a higher
supporting electrolyte concentration than is sufficient for
steady state.
Dickinson’s[104] results are summarised in Figure 17, which
shows simulated cyclic voltammograms for the reduction of
some neutral species A at a 1 mm radius hemispherical elec- Figure 18. Simulated cyclic voltammograms for a fully reversible one-elec-
trode and a scan rate of 0.5 V s¢1 at various levels of support. tron reduction in water at a hemispherical electrode for various charges of
electroactive species. Parameters: cA = 1 mm, DA = 1 Õ 10¢8 m2 s¢1, re = 25 mm,
The zero field approximation described briefly above was used
n = 1 mV s¢1, csup = 1 mm. All other diffusion coefficients are equal to DA.

electrical migration towards the electrode. Conversely, the neg-


atively charged species has a smaller limiting flux since it is re-
pelled away from the build-up of negative charge around the
electrode. It is worth noting that if the electroactive species is
charged, then it and its counter ion can act as supporting elec-
trolyte, resulting in a “self-supported” system. The effect of an-
alyte charge is shown in Figure 18. This figure shows cyclic vol-
tammograms for a fully reversible reduction of some species
with a charge of 0, + 1, and ¢1 at a 25 mm radius hemispheri-
cal electrode and a scan rate of 1 mV s¢1, with 1 mm monova-
lent supporting electrolyte present.

5.4 Applications of weakly supported voltammetry


Figure 17. Simulated cyclic voltammograms for a fully reversible one-elec-
As alluded to above, the effects of migration in weakly sup-
tron reduction in water at a hemispherical electrode at various support
ratios. Parameters: CA = 1 mm, DA = 1 Õ 10¢9 m2 s¢1, re = 1 mm, v = 0.5 V s¢1. ported voltammetry may be used to extract valuable kinetic
All other diffusion coefficients are equal to DA. and mechanistic data unobtainable if experiments are carried

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 246 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
out under full support. Three examples of this are discussed 2.5 Õ 107 dm3 mol¢1 s¢1 to 2.5 Õ 109 dm3 mol¢1 s¢1.[139] Barnes
below. In general the lack of supporting electrolyte leads to et al.[129b] investigated this reduction at a range of concentra-
the cyclic voltammetry becoming sensitive to the charge of tions of tetrabutylammonium tetrafluoroborate supporting
the reactants, intermediates, and product species. electrolyte, and using simulation, were unable to reproduce ex-
perimental data across the whole range of supporting electro-
lyte concentration used. It was found that the experimental
5.4.1 Comproportionation in the reduction of anthraquinone peak heights, instead of leveling off at high support (as in
If two successive electron transfers occur as in the following Figure 17), continued to decrease (the peak height increased
mechanism: with reducing concentration of supporting electrolyte due to
migration of the positive DPP + ion; see above).
To account for this, a further mechanistic step was recog-
A þ e¢ Ð B E f1 ð55Þ
nised: a fast ion-pairing equilibrium of the DPP + ion and the
tetrafluoroborate anion, BF4¢ :
B þ e¢ Ð C E f2 ð56Þ

then, providing E f1 > E f2 Ef1 > Ef2 for reductions as shown (or, if kf
DPPþ þ BF¢4 Ð X KIP ¼ ð60Þ
the electron transfers were oxidations, E f1 < E f2) then compro- kb
portionation between species A and C becomes thermody-
namically (but not necessarily kinetically) favourable: where KIP (dm3 mol¢1) is the equilibrium constant, defined as
the ratio of forward and reverse rate constants, kf
A þ C ! 2B kcomp ð57Þ (dm3 mol¢1 s¢1) and kb (s¢1) respectively. With this step included
in the mechanism, experimental voltammetry was successfully
Andrieux and Sav¦ant[138] showed that, under diffusion only simulated across the whole range of supporting electrolyte
conditions, if both electron transfers are fully reversible and all concentrations and scan rates used. The concentration of free
diffusion coefficients are equal, then cyclic voltammetry will be DPP + was decreased when a large amount of supporting elec-
completely insensitive to the presence of comproportionation. trolyte was present, and the peak currents therefore decreased.
Hence diffusion only cyclic voltammetry will be inadequate to This resolved the literature controversy over the rate of dimeri-
establish whether or not comproportionation takes place. sation, and a value of kdim = 5 Õ 105 dm3 mol¢1 s¢1 was estab-
Belding et al.[129d] investigated the stepwise two-electron re- lished, with an equibibrium constant for ion pairing of KIP =
duction of anthraquinone in nonaqueous solvents at both 35 dm3 mol¢1. These observations were only made possible by
high and low concentrations of supporting electrolyte to de- the ability to simulate cyclic votlammetry at varying concentra-
termine if the added effects of migration would allow the pre- tions of supporting electrolyte.
sense or absence of comproportionation to be determined. Ex-
perimental data collected at high support was simulated, and
good agreement was seen. Simulations were carried out both
in the absence of comproportionation and in the presence of 5.4.3 The ECE/DISP1 mechanism
fast comproportionation. The same experiments and simula- A mechanism often encountered in electrochemistry, especially
tions were carried out at low concentrations of supporting in the reduction of aromatic halides, is the ECE/DISP1 mecha-
electrolyte. When comparing simulated data in the presence nism:[140]
and absence of comproportionation, differences were seen,
with the second reductive peak significantly reduced in size A þ e¢ Ð B¢ E f3 ð61Þ
when comproportionation was fast. Comparison of these simu-
lations to experimental data confirmed the presence of fast B¢ ! C ¢ k C ð62Þ
(kcomp > 108 dm3 mol¢1 s¢1) comproportionation between anthra-
quinone and its dianion, which is impossible to detect using C¢ þ e¢ Ð P2¢ E f4 ð63Þ
conventional diffusion-only voltammetry.
B¢ þ C¢ Ð A þ P2¢ kdisp ð64Þ

5.4.2 Ion pairing in electrochemical mechanisms If Ef4 > Ef3 for the reductions as written, then the disproporti-
nation step 64 is thermodynamically viable. The source of the
It is known[22, 139] that the reduction of the diphenylpyrylium
second electron transfer may then be either step 63 (the ECE
(DPP) cation follows an EC2 type electrochemical mechanism,
mechanism) or step 64 (the DISP1 mechanism if step 62 is rate
with ion pairing of the DPP radical occurring after reduction:
limiting, or the DISP2 mechanism if step 64 is rate limiting).
DPPþ þ e¢ ! DPP ð58Þ While DISP2 is able to be distinguished from the other mecha-
nisms with relative ease,[141] the discrimination between ECE
2DPP ! DPP2 kdim ð59Þ and DISP1 is difficult.[142] At high concentrations of supporting
electrolyte, careful analysis of voltammetric wave shape over
The literature reports a large range of values for kdim from a range of scan rates can in some circumstances be used.[129c]

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 247 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
At steady state, however, it is impossible to distinguish the where the net rate of the electrochemical process is given by
two mechanisms.
If, however, migration is included as a mode of mass trans-
j ¼ k0 ½expð¢aqÞ½ A¤0 ¢ expðð1 ¢ aÞqÞ½B¤0 ¤ ð66Þ
port through addition of only a small amount of supporting
electrolyte, then discrimination may become much easier
and
through analysis of peak heights as a function of supporting
electrolyte concentration, as shown by Barnes et al.[129a]  Ž
If the chemical step 62 is such that species B and C have the q ¼ F E ¢ Efo;A=B =RT ð67Þ
same charge, as written above (for example, an isomerisation),
then species A and C, the two possible sources of the second k0 is the standard electrochemical rate constant, a is the
electron in the mechanism, will undergo migration to different Butler–Volmer transfer coefficient[9, 143] and [X]0 is the surface
extents (as written in the mechanism above, species A will not concentration of species X. It is well known that as k0 decreas-
migrate at all, and species C will migrate away from the elec- es in size, an overpotential is required to ‘drive’ the electrode
trode as negative charge is introduced via reduction). This process. In terms of cyclic voltammetry this is revealed by an
means that if the mechanism is ECE, and species C is the increase in the peak-to-peak separation in terms of the poten-
source of the second electron, the peak current will be re- tial as shown in Figure 19 which has been calculated for a typi-
duced, since species C is repelled from the electrode. If the cal macroelectrode, radius 0.15 cm, and a voltage scan rate of
mechanism is DISP1, and species A is the source of the second
electron, migration will not reduce the current since species A
is electrically neutral. It is therefore possible, in theory, to
assign the mechanism as ECE or DISP1 by carrying out experi-
ments in the presence of both high and low concentrations of
supporting electrolyte, and simulating both cases. Only one
mechanism should give a consistent value of kc over the whole
range of supporting electrolyte concentrations used.
Understanding of weakly supported voltammetry has been
greatly enhanced in recent years through the use of numerical
simulation. Such simulations have been used not only to gain
a more fundamental insight into electron transfer processes
occurring inside a double layer, but have also been applied to
model experimental data. The latter in particular has allowed
for greater insight into chemical processes and the discovery
of new, sometimes unexpected, information. This demonstrates
the power of this technique.
Figure 19. Simulation of one-electron oxidation process. a = b = 0.5,
DA = DB = 1 Õ 10¢5 cm2s¢1, Eo = 0, n = 0.1 V s¢1, electrode area = 0.0707 cm2,
C*A = 0.001 m, k0 ranges from 1 cm s¢1 to 1 Õ 10¢10 cm s¢1. Reproduced with
Section 6: Voltammetry at Rough and Porous permission from Ref. [143]. Copyright 2010, Elsevier.
Surfaces
0.1 V s¢1. The value of k0 ranges from 10¢12 to 10¢2 m s¢1, which
Much current experimental electrochemistry is a materials-
spans the range from electrochemically irreversible, through
based activity devoted to the search for electrocatalysts that
quasi-reversible to fully electrochemically reversible, where the
might assist various technologically important electrode pro-
term electrochemical reversibility indicates the speed of the
cesses including, for example, the reduction of oxygen, the oxi-
electron transfer (k0) relative to the prevailing rate of mass
dation of methanol, the oxidation of formic acid, and the hy-
transport (kMT ~ D/r where D is the analyte diffusion coefficient).
drogen evolution reaction. Voltammetric methods are widely
The Randles–Ševč†k equation for the voltammetric peak cur-
employed to test the success of the electrocatalysts, generally
rent of a reversible overall n-electron reduction process is:
deployed so as to modify the electrode surface, even though
the nature of the cyclic voltammetric experiment is rather dif-
ferent from the conditions in which the catalyst is likely to be Ip;rev ¼ 2:69 105 ðpr2 Þn3=2 Cbulk D1=2 v1=2 ð68Þ
employed in, say, a fuel cell or battery.
The term ‘catalyst’ implies a change in the rate constant for whilst that of an irreversible process is:
the process of interest either via a change of mechanism or via
the lowering of the activation energy within the same mecha-
Ip;irrev ¼ 2:99 105 ðpr 2 Þðn0 þ aÞ1=2 nCbulk D1=2 v1=2 ð69Þ
nism. To explore the role of an electrochemical rate constant,
we consider a simple one-electrode process
where n’ is the number of electrons transferred before the
rate-determining step (for which the electron transfer is char-
A þ e¢ Ð B ð65Þ acterised by a).

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 248 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
It is interesting to calculate the ratio of the peak currents for porous layer and, second, a component due to diffusion trans-
reversible and irreversible processes. For the case of n = 1, a = port within the porous layer.[144] If the packing density and
b = 0.5 and n’ = 0, thickness of the porous layer is suitably large (but typical) then
the transport within the porous layer can be approximated as
Ip;rev a ‘thin layer’ in which the distance diffused by the solute in
¼ 1:27 ð70Þ order to reach a location where it can be electrolyzed is short
Ip;irrev
as compared to that which occurs under semi-infinite diffusion
as can be seen from Figure 19[143] described above. It follows conditions, and hence, the (diffusional) overpotential is mark-
that changing the electrochemical rate constant in this case edly reduced reaching, in the limit, apparent reversible ‘thin-
has a rather tiny effect on the peak current; the effects of layer’ behavior. Figure 20 shows the effect: it compares cyclic
changed k0 (and hence electrocatalysis) are best judged by the
peak-to-peak separation, although if the reaction is chemically
irreversible, and the processes for which electrocatalyts are
most sought are typically of this type, then the absolute po-
tential of a single peak potential might be used but subject to
the caveats raised below.

The rather low value of a 27 % decrease in current between


the electrochemically reversible and irreversible limits relates
to the case of n = 1. For the case of an n > 1 process, and as-
suming that the rate-determining step under irreversible con-
dition is the first electron transfer,

Ip;rev 2:69 n3=2  n Ž1=2


¼ ¼ 0:9 ð71Þ
Ip;irrev 2:99na 1=2 a

This is larger (than 27 %) but still relatively modest.


Figure 20. Comparison of linear sweep voltammetry using a semi-infinite
The insensitivity of the peak current to the rate of electron and thin layer planar diffusion models. For both models, k0 = 10¢4 cm s¢1;
transfer reflects the fact that the voltammetric peak arises as D = 10¢5 cm2 s¢1; n = 0.1 V s¢1; c = 10¢6 mol cm¢3. Semi-infinite diffusion elec-
a competition between the two processes of mass transport trode area, A = 1 cm2 ; thin-layer area, A = 30 cm2 ; thickness, l = 1 mm. Repro-
duced with permission from Ref[144a]. Copyright 2008, Elsevier.
(diffusion) and electron transfer. The calculation above as-
sumed semi-infinite diffusion-only conditions, linear diffusion,
and a flat and planar macroelectrode. In the case of conver-
gent diffusion to a microelectrode, the enhancement of the
current, of course, will be less since under true steady-state voltammetry for a thin-layer system (area 30 cm2, thickness
conditions, a limiting current will flow, reflecting simply the 1 mm) with that seen for semi-infinite diffusion from a system
total number of electrons transferred: with D = 10¢5 cm2 s¢1 and Cbulk = 1 mm for a standard electro-
chemical rate constant of 10¢4 cm s¢1 (and a = 0.5). Note the
Ilim ¼ 4nFr D Cbulk ð72Þ much reduced peak-to-peak separation for the thin-layer case
with the contrast from semi-infinite diffusion arising solely
It is important to next explore what happens if the electro- from the altered mass transport.
de’s surface remains that of a macroelectrode but is ‘modified’ The shift in the peak potential under increasingly thin-layer
with a layer of catalyst so as to create a porous structure on conditions will ultimately (as the distance required for diffusion
the electrode surface, reflecting many of the electrode modifi- is steadily reduced) tend to a limiting value where the poten-
cation strategies used in energy research. tial corresponds to that of the formal potential of the couple
of interest. This observation explains the fact that the modifica-
tion of an electrode with an electrochemically conductive
6.1 Semi-infinite diffusion versus thin layer: qualitative
porous layer can resolve otherwise overlapping voltammetric
insights
peaks. This has potential analytical value[144b] and can simply
A particular important case is when a macroelectrode is modi- arise as a consequence of the altered mass transport; it is not
fied with a conductive porous layer made, for example, of necessary to invoke changed electron transfer kinetics.
carbon nanotubes, graphene, or nanoplatelets. In this situation, Finally we point out that under extreme conditions it should
the mass transport of solute to the conductive surface (as- be noted that it is possible to see two peaks resulting from
sumed to be both the substrate electrode and the conductive, a single A/B redox couple: one is a thin-layer signal and the
modifying layer) arises from two components: First, semi-infin- other that arising from semi-infinite diffusion to the surface of
ite diffusion from the solution bulk to the surface of the the porous layer.[144d]

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 249 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6.2 Carbon nanotube and other modified electrodes

The effects predicted for thin-layer versus semi-infinite diffu-


sion voltammetry are consistent with observations made using
a range of systems.[144a–c] In particular, there has been very con-
siderable work in using carbon nanotubes (or chemically modi-
fied nanotubes) to create porous layers on the surface of elec-
trodes. The observed voltammetry is consistent with the ideas
outlined in the previous section (refs. [144a–c] and refs. there-
in). Figure 21 shows the basic model in which the trapped

Figure 22. Electrode geometry types discussed in Section 6.3. Shaded re-
gions are electroactive, white areas are inactive. Reproduced with permis-
sion from Ref. [145b] Copyright 2008, Elsevier.

Figure 21. Schematic of the two types of diffusion that contribute to current
at a carbon-nanotube-modified electrode. Reproduced with permission from
Ref. [144a]. Copyright 2008, Elsevier.

products of analyte-containing solution act as small thin-layer


cells. A key indicator of this behavior is the observation that
the peak currents flowing associated with the thin-layer behav-
ior can be significantly larger than those seen for semi-infinite
diffusion at an electrode of the same geometric area. Note, as
discussed above, such large enhancements are not under-
standable in terms of altered electrode kinetics per se. The
work of Henstridge et al.[144b] contains tables of examples of Figure 23. A) Schematic of a porous surface; B) side on view. Reproduced
CNT and other modified electrodes in which the thin-layer be- with permission from Ref. [145a]. Copyright 2014, Elsevier.
havior may operate. Similarly Kozub et al.[143] report ‘electroca-
talytic’ systems developed allegedly for the detection of nitrite. " rffiffiffiffiffiffiffiffiffiffiffi#
RT aFDv
Ep ¼ Efo ¢ 0:780 ¢ lnðYko Þ þ ln ð74Þ
aF RT

6.3 Effective heterogeneous rate constants for rough and for a one-electron process. Note that this relationship assumes
porous surfaces that the adjacent particles are sufficiently close together, that
on the timescale of the experiment the diffusion field is
Numerical simulation[145] has been explored to identify the ef-
normal to the place of the bulk electrode surface; that is to
fective standard electrochemical rate constant for both rough
say, the diffusion fields of the particles are heavily overlapped,
and porous surfaces using surface morphologies such as those
which is usually the case except for very dilute coverages.
shown in Figure 22 and 23. The former models a dense, but
less than monolayer array of nanoparticles whilst the latter ap- In the case of a porous surface,[145a]
proximates a porous surface. In the former case, the apparent
electrochemical rate constant (inferred from the peak potential ze
Y ¼ 1 þ 2V ð75Þ
of the voltammetry) was seen to vary according to re

o where ze is the depth of the pores, re is their radius and V is


kapp ¼ Yko ð73Þ
the porosity defined by,

where ko is the true electrochemical rate constant, and Y is


the ratio of the electroactive surface area to the geometric sur- re2
V¼ ð76Þ
face area.[145b] In this case the peak potential is given by rd2

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 250 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 24. Unit cell and coordinate system for a porous surface with coordi-
nates r, z, and f. Reproduced with permission from Ref. [145a]. Copyright Figure 25. Schematic design of the thin-layer model used in this article,
2014, Elsevier. where x is the distance perpendicular to the electrode surface (x = 0). The
film thickness is xf. Reproduced with permission from Ref. [148]. Copyright
2014, PCCP Owner Societies.
where re and rd are shown in Figure 24 as defined by the diffu-
sion domain approximation to the electrode surface.[145] The ef-
Figure 25, in which the modifying layer itself is nonelectroac-
fects of the roughness and porosity are illustrated in the fol-
tive but changes the voltammetric response by virtue of alter-
lowing subsection but an important general extension relates
ing the solubilites and diffusion coefficients of the electroactive
to the use of rotating disc electrodes to extract the number of
species within the layer as compared to the bulk solution. The
electrons (n) transferred and the electrochemical rate constant
electron transfer in this scheme is limited to the surface of the
by measuring the transport limited current as a function of
substrate electrode. Both cyclic voltammetry[148] and electro-
disc rotation speed and applying the Koutecky–Levich equa-
chemical impedance spectroscopy (EIS)[149] were modelled; it
tion.[146] It was found that this analysis gives correct values of n
was established that the illusion of altered electron-transfer
but the apparent, rather than true, rate constant as defined
characteristics could be generated in each case merely by al-
above.[146]
tered diffusion or solubility. The difficulty of unambiguously
modelling the response was noted.
6.4 Evaluating oxygen reduction catalysts
Section 7: Nanoparticle Voltammetry
The results of the preceding section have been applied in par-
ticular to a consideration of nanoparticle-modified electrodes As discussed in the previous section, alteration of the electrode
for oxygen reduction catalysis.[147] Such evaluations often in- surface structure can lead to apparent changes in the electron
volve simply a measurement of a current at a fixed potential. transfer kinetics of an electrochemical reaction; this arises due
Simulations of the type reported in the previous section clari- to changes in the mass-transport regime local to the interface,
fied that such currents are sensitive to the surface coverage of and alters what would classically be referred to as the ‘diffu-
nanoparticles without any change in the fundamental kinetics sional overpotential’.[150] In light of this insight, this section fo-
or thermodynamic parameters, even if the voltammetry shows cuses on understanding the voltammetry of electrodes modi-
that the reduction operates under full diffusional transport fied with submonolayer coverages of nanoparticles, how true
control. The need for caution in the evaluation of catalysts in electron transfer rates may be extracted via simulation from
the manner discussed was evident and an essential need for the experimental data, and how the diffusion field influences
characterising coverage, porosity, and particle size demonstrat- the stripping voltammetry of nanoparticles. Using the strategy
ed for establishing authentic electrocatalytic character. outlined within this section, it is now possible to rigorously in-
vestigate the presence or absence of ‘nano-effects’ arising
from the use of novel nanomaterials in voltammetric experi-
6.5 Voltammetry at thin-layer, nanoparticle-modified elec-
ments. In the final part of this section, an alternative technique
trodes
for studying nanoparticle electrochemistry is highlighted. This
The previous section concerned electrodes modified with elec- new technique referred to as ‘nano-impacts’ exhibits a number
trochemically conductive layers or films. A more general situa- of advantages over more conventional electrochemical investi-
tion has been modelled using the scheme summarised in gative techniques.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 251 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
7.1 Reactions on nanoparticle-modified surfaces Ilim ¼ 4pnFDCr ln 2 ð77Þ

For an array of nanoparticles supported upon an electrochemi- At longer experimental times, the diffusion layers of adjacent
cally inert electrode substrate, the mass transport to and from particles will overlap strongly. Under such situations, classified
the nanoparticulate surface, and hence the voltammetric be- as case 4, the mass transport to the whole electrode surface is
haviour of the electrode, depends upon: the nanoparticle’s linear, resulting in a macroelectrode response with an associat-
size and morphology, the diffusion coefficient of the analyte ed apparent electrochemical rate constant. The magnitude of
and product, the experimental (voltammetric) time scale, and this rate constant is a function of the nanoparticle surface cov-
the interparticle separation (nanoparticle surface coverage).[151] erage.[145b] Transition between isolated nanoparticle diffusion
Assuming the substrate electrode is macroscopic in dimen- layers (case 2) and strongly-overlapping layers (case 4) leads to
sions, then the diffusion regime may be categorised into four case 3. Case 3 is a situation regularly encountered with modi-
cases. Importantly, during the course of a voltammetric scan, fied nanoparticle surfaces where the diffusional layers are par-
the prevailing diffusion regime will likely transit from one case tially overlapping and can only be approached through simula-
to another; consequently, insight into the voltammetry of such tion. A fifth diffusional case may also be considered as an ex-
systems is best achieved through simulation. Figure 26 sche- tension of the above model; this case arises for the situation in
matically outlines the four diffusional cases or categories. which the dimensions of the supporting electrode is only of
the order of microns.[153] Under this case, the mass transport to
the whole array is convergent. This category may be viewed
partially as a breakdown of the diffusion domain approxima-
tion used in macroelectrode simulation models and is notably
of significance in application to some SECM experiments.[154]
Due to the complex interplay between the electrode surface
geometry and the diffusion profile as outlined above, extrac-
tion of physically significant kinetic data from modified electro-
des is a nonfacile problem. To experimentally evidence this
point, the one electron reduction of chromium(III)[155] has been
studied at a silver-nanoparticle-modified electrode and a silver
macroelectrode, demonstrating clearly how the voltammetric
response varies as a function of nanoparticle surface coverage.
It should also be commented that normalisation of voltammet-
ric results relative to the total electroactive area is also insuffi-
cient to, in many cases, allow even a qualitative comparison of
data.[145b] Theoretically this conclusion has been specifically va-
lidated in relation to the oxygen reduction reaction at nano-
particle-modified surfaces.[147] Moreover, experimentally, it has
been demonstrated how for sparse coverages of platinum
Figure 26. Simulated concentration profiles at a diffusion domain containing nanoparticles on an electrochemical interface, the situation is
a spherical particle. Category 1: s = 1000. Category 2: s = 10. Category 3: further complicated by the release of the hydrogen peroxide
s = 1. Category 4: s = 0.01, where s is the dimensionless scan rate. Concen- as an intermediate in the reduction process.[156] As discussed
tration profiles were taken at the linear sweep’s peak potential. Reproduced
within Section 1, this alteration of the electrochemical mecha-
with permission from Ref. [151b]. Copyright 2007, American Chemical Soci-
ety. nism, arising from the enhancement of the mass transport to
and from diffusionally isolated nanoparticles has large implica-
tions not just for the industrial use of such nanomaterials as
First, at very short times, the diffusion profile at each individ- catalysts, but may also be of importance in the context of the
ual nanoparticle is linear; this situation occurs for times of the toxicity of these substances within biological systems.
order of rNP2/D (commonly ~ 1 ms, where D ~ 10¢10 m2 s¢1 and Figure 27 outlines a general strategy for the combined ex-
rNP ~ 10 nm). As the experimental time increases, the diffusion perimental and computation study of electrocatalytic process-
layer grows in accordance with the Einstein equation (d ~ es at electrodes modified with ensembles of nanoparticles.
(2 Dt)0.5); overlap between the diffusion layers of adjacent This methodology allows the influence of the mass transport
nanoparticles occurs when the diffusion layer thickness d is and the interfacial electron-transfer kinetics upon the voltam-
comparable to the interparticle separation. Case 2 arises for sit- metry to be clearly delineated.[157] Briefly, the approach requires
uations where the nanoparticle diffusion layers do not signifi- characterisation of the electrode in terms of particle size, ag-
cantly overlap, but the experimental time is greater than gregation, and separation, allowing the voltammetric response
rNP2/D. Here, the mass transport to each nanoparticle is conver- to be simulated using the kinetic parameters obtained using
gent and can be considered independently of adjacent parti- a macroelectrode. From comparison of the simulated ‘bulk’ re-
cles. For an isolated sphere on a surface, the steady-state diffu- sponse and the experimentally recorded data, it is possible to
sion-limited current is given by;[152] determine if the kinetics have been altered while fully account-

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 252 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
and the hydrogen evolution reaction, where for small gold
nanoparticles (1.9 nm diameter), the electrochemical processes
were found to be significantly hindered as compared to the ki-
netics recorded on the macroelectrode.[158] This is a prime ex-
ample of how decreasing the size of the nanoparticles has led
to a ‘negative’ electrocatalytic effect, likely resulting from the
changed reaction intermediate adsorption on the gold surface.

7.2 Stripping voltammetry: the direct oxidation and reduc-


tion of nanoparticles
The above examples focus on the situation in which the nano-
particle-modified electrode is utilised to study the electro-cata-
lytic properties of the nanoparticles towards a given redox pro-
cess. However, the diffusional cases outlined above must also
be considered when studying the direct oxidation or reduction
of the nanoparticles themselves—as is commonly undertaken
during the course of an electrochemical nanoparticle stripping
experiment. Such nanoparticle stripping experiments are pur-
portedly a facile route to directly study the possibly altered
thermodynamics of the metallic nanoparticles[159] and their in-
teractions with the electrode substrates.[160] However, problems
arising in such experimental setups relating to partial oxidation
or reduction of the nanomaterial has been previously noted—
a problem that is circumvented by the use of nano-impacts ex-
Figure 27. Flow diagram showing the proposed strategy for delineating
periments (see final part of this section for further discus-
kinetic and mass-transport effects allowing the detection of authentic nano-
electrocatalysis. sion).[161]
In nanoparticle stripping voltammetry, for the case in which
the product of the nanoparticle oxidation or reduction is solu-
ble, the diffusion layer associated with the redox product must
ing for the diffusional mass transport of the material. From this be considered. Both simulations[162] and analytical expres-
analysis three outcomes are possible. If the nanoparticle array sions[163] have been previously provided demonstrating clearly
simulation using the ‘bulk’ kinetics is in good agreement with how, for the stripping of nanoparticles from an electrode sur-
the experimentally recorded data for the nanoparticle array, face, the observed peak potential varies as a function of the
then the conclusion must be that there is no evidence of total amount of electroactive material on the electrode surface,
a nano-effect associated with using the nanomaterial. Con- the inter-particle distance and the voltammetric scan rate.
versely, if the simulated voltammetric response differs from Moreover, due to the sensitivity of voltammetry towards the
that found experimentally, then the kinetics of the electro- mass-transport regime local to the interface, stripping voltam-
chemical reaction must have been altered through the use of metry can also be used as a route to indirectly evidence ag-
the nanomaterial. This alteration in the kinetics may lead to an glomeration of the nanoparticles upon the electrode sur-
increase in the overpotential required to drive the reaction; face.[163] In accord with nanoparticle electrocatalysis studies,
hence, for such situations, the nanomaterial is less catalytic care must be taken in the analysis of the voltammetric re-
than the bulk material, and one has evidenced a ‘negative’ sponse, accounting for the influence of diffusion prior to as-
nano-effect. Alternatively, if the overpotential required for the cribing the altered peak positions as relating to changes in the
electrochemical reaction is decreased as compared to the simu- thermodynamics of the nanoparticles. Finally, the use of micro-
lated result then it can be confirmed that the use of the nano- electrodes in stripping voltammetry (an analogous diffusion
material has led to an authentic nano-effect. This procedure regime to Case 5 discussed above) may provide one route by
was applied to the experimental study of arrays of gold nano- which true nanoparticle thermodynamic effects may be more
particles ranging in size from 20 to 90 nm in diameter. From readily investigated, aided by the relatively well-defined
these experiments, it was confirmed that for nitrite electro-oxi- steady-state diffusion.[164]
dation, no alteration in the kinetics is observed between the Figure 28 depicts the stripping voltammetry for two sizes of
use of gold nanoparticles and a gold macroelectrode.[157] Con- nanoparticles at two different surface coverages from
versely, the electro-oxidation of l-ascorbate was shown to ex- a carbon-fibre microelectrode (r0 = 5.5 mm), where the ob-
hibit true nanocatalytic effects. The origins of these differences served shift is consistent with a change in the thermodynamics
were ascribed as likely being due to the l-ascorbate oxidation due to the influence of the altered surface energy. Note that
involving adsorbed intermediates.[157] Finally, this same meth- the observed peak potential also varies in both cases as a func-
odology has been applied to the oxygen reduction reaction tion of the total surface coverage of silver. Not all nanoparticle

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 253 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 28. Anodic stripping voltammetry for small (rnp = 3.7 nm, dashed line) and large (rnp = 13.5 nm, solid line) silver nanoparticles supported on a micro
carbon-fibre electrode (r0 = 5.5 mm): A) low surface coverage of silver (1.6 Õ 105 mol m¢2) and B) high surface coverage (1.6 Õ 104 mol m¢2). Scan rate: 50 mV s¢1
and supporting electrolyte of 0.1 m NaClO4. Reproduced with permission from Ref. [164]. Copyright 2014, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

redox reactions result in the dissolution of the nanoparticle ‘nano-impact’ experiments as will be discussed in the next and
(though there will be a corresponding change in the nanopar- final section.
ticle morphology). One example of such a case is the oxidation
of silver in the presence of a halide. Depending on the halide
7.3 ‘Nano-impacts’
concentration, the electrochemical oxidation will likely lead to
the formation of surface bound silver halide.[165] A secondary In ‘nano-impact’ experiments the nanoparticles are suspended
example would be the electrochemical formation of metal within an electrolyte and randomly, by virtue of Brownian
oxides.[166] The voltammetric response of such systems is highly motion, collide with a potentiostated microelectrode.[170] Upon
complicated; first and foremost, the formation (or solubility) impact the nanoparticle makes electrical contact, and assum-
constant between the formed nanoparticle ion and the solu- ing a suitable potential is held upon the electrode, either the
tion-phase counter-ion serves to alter the thermodynamics of direct electrochemistry of the nanoparticle or a catalytic pro-
the redox species in accordance with the Nernst equation.[167] cess of interest may be induced. The resulting current yields
Second, the reduction or oxidation may exhibit complex be- direct and significant information regarding the interfacial
haviour such as following a nucleation growth mechanism.[168] redox processes occurring at individual particles.
Third, speciation of the products may vary as a function of The use of ‘mediated’ nano-impacts has been demonstrated
counter-ion concentration.[169] Finally, in some cases the reac- for a variety of systems.[171] One important factor in such ex-
tion may be limited by the mass transport of the counter-ion periments is that the electrode receiving the nanoparticles is
to the electrochemical interface. Consequently, again when inert, such that the reaction of study does not occur on its sur-
studying these systems, care must be taken when ascribing face in the absence of the nanoparticle. This requirement has
any alteration in the stripping peak potentials as relating to al- led some researchers to use mercury as the ‘inert’ electrode
tered thermodynamics or ‘nano-effects.’ This is especially true substrate.[172] However, problematically in the presence of trace
due to the fact that the nanoparticle capping agents have chloride, mercury is readily oxidised to form calomel nanoparti-
been shown to influence the observed stripping voltamme- cles.[173] Once formed, these calomel nanoparticles are easily re-
try.[161] As an interesting aside, the strength of the binding of duced at the mercury electrode and can lead to results which
silver ions to halides and the corresponding Nernstian shift in may be misinterpreted. Consequently, in many cases, carbon
the stripping peak yield an analytically useful route to their de- electrodes are found to be more suitable candidates for the
tection via the use of nanoparticle stripping voltammetry.[167] supporting electrode material. A notable exception to this is
From the above discussion in the last two sections, it is clear an early example of the use of nano-impacts for study of the
that although not unsurmountable, the interpretation and oxidation of hydrazine on platinum nanoparticles at a gold
measurement of physically significant values from the electro- electrode.[174] The difference in the electron-transfer rate on
chemical response of electrodes modified with ensembles of platinum versus gold is sufficient to allow the process to be
nanoparticles is inherently challenging. Consequently, there is solely studied upon the impacting platinum nanoparticle. It
a desire to find methods by which nanoparticles and their should, however, be noted that for this system, the presence
properties can be studied individually. One route through of hydrazine in solution does cause significant aggregation of
which this has been achieved is with the use of so-called the particles.[175]

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 254 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
One significant observation for impact experiments is that served in the current during the course of a nanoparticle
the recorded collision frequency is not uncommonly below impact was interpreted as likely indicating a contact resistance
that theoretically predicted.[171b] This observation has previous- between the electrode and the impacting nanoparticle. Conse-
ly been explained solely in terms of solution-phase agglomera- quently, although the presence of the nanoparticles in solution
tion/aggregation.[175] However, other causes for such effects can be evidenced from their catalytic impacts at an electrode,
need to be considered. In many experimental cases the work- the full potential of this method has yet to be realised.
ing electrodes used tend to be micron sized wires sealed in Apart from the study of the catalytic response of impacting
glass. Consequently, if the nanoparticles of study also adhere nanoparticles, in analogy with nanoparticle stripping voltam-
to the inert (glass) substrate surrounding the electrode, then metry, the direct redox of the impacting nanoparticle may also
the frequency of observed nanoparticle impacts may be signifi- be studied. The first example of such an experiment was the in
cantly reduced due to diffusional shielding.[176] A further issue situ detection of silver nanoparticles.[180] Upon impacting an
arises upon consideration of the dimensions of nanoparticle as electrode with a suitably anodic potential, the nanoparticles
compared to a molecule. As a result of the geometrical con- were oxidized, resulting in small spikes of current. Through
straints of an impacting nanoparticle, the ‘nano-impact’ experi- Faraday’s first law, the magnitude of these spikes can be relat-
ment will be inherently more sensitive to the presence of ad- ed directly to the total number of atoms contained within the
sorbing and blocking organic media than conventional molec- individual impacting nanoparticle, hence providing measure-
ular redox probes.[177] Figure 29 depicts a simple geometric ment of its size.[161, 181] Furthermore, the frequency of the im-
pacting spikes can yield information regarding the concentra-
tion of the nanoparticles in solution.[182] This new nano-metrol-
ogy technique has been successfully applied to a host of differ-
ent nanoparticle materials including other metals,[183] metal
oxides,[184] carbon fullerenes,[185] organic nanoparticles,[186] and
liposomes.[187] Work has also demonstrated that the technique
is suitable for use in complex media such as sea water, high-
lighting its potential use in environmental analysis.[188] To this
end the use of carbon-fibre microcylinder electrodes has been
advocated as one route by which subpicomolar concentrations
of nanoparticles may be readily detected and sized,[189] the sec-
ondary advantage of this methodology is the minimisation of
problems associated with nanoparticle adsorption onto the
electrode support.
Figure 29. Diagram for the calculation of the minimal distance, dmin, to Beyond being an analytical technique, the direct nano-
a blocking molecule, at which a nanoparticle can touch the electrode sur- impact method also provides a route for investigating more
face (impact). Reproduced with permission from Ref. [177] Copyright 2014,
fundamental problems. A first example is the use of the tech-
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
nique for the study and monitoring of solution-phase nanopar-
ticle agglomeration and aggregation.[190] In a similar vein, the
model showing how the presence of surface-adsorbed species magnetic-field-induced agglomeration of Fe3O4 has also been
leads to efficient blocking of the electrode surface towards directly evidenced.[191] Figure 30 depicts examples of iron oxide
nanoparticles due to the magnitude of the minimum distance reduction spikes and the associated size distributions obtained
(dmin) of an impacting nanoparticle to the blocking molecule. in the presence and absence of a magnetic field. Second, the
The magnitude of this minimum distance of approach depends electron-transfer kinetics and mechanism to the nanoparticles
upon both the radius of the nanoparticle (rp) and the height of can be studied.[183b, 192] Third, the interaction of the nanoparticle
the adsorbed species (hb). Hence, the observed decreased with the electrochemical double layer is of utmost importance
impact frequency may not just be related to agglomeration or and the nano-impacts methodology has provided a direct
aggregation of the nanoparticles in solution, but may also route by which these interactions can be probed.[193]
arise due to factors relating to the electrode and the electrode Consequently, we conclude that the use of ‘mediated’ and
design itself. ‘direct’ nano-impact experiments have significant potential for
Although each individual impact experiment (chronoam- future research both as nano-metrology techniques and as
perogram) is studied at a fixed potential, performing repeat ex- novel methods for addressing fundamental and technological
periments at differing potentials can yield significant informa- nanoparticle challenges.
tion regarding the kinetics of the catalytic process on the
nanoparticle.[178] Comparison of the kinetics for the reduction
Conclusion
of protons at an impacting gold nanoparticle compared to
that found from more conventional ensemble measurements The above survey shows that with the era of numerical simula-
led to the conclusion that the process was significantly slower tion, the technique of voltammetry has come of age and has
on the individual impacting gold nanoparticle.[179] This appar- power to contribute very significantly, not least in a quantita-
ent decrease in the kinetics and the notable fluctuations ob- tive manner, to many problems in analytical, physical, and bio-

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 255 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Figure 30. Chronoamperograms showing reductive impact spikes recorded without (A) and in (B) a magnetic field and the derived impact charge and NP size
distributions (C and D); electrolyte: 0.2 m phosphate buffer (pH 10), E (vs. SCE) = 0.9 V. Reproduced with permission from Ref. [191] Copyright 2014, PCCP
Owner Societies.

[1] a) M. Rudolph, J. Electroanal. Chem. 1990, 292, 1 – 7; b) M. Rudolph, J.


physical chemistry, as well as to nanochemistry and nanotech- Electroanal. Chem. 1991, 314, 13 – 22; c) M. Rudolph, J. Electroanal.
nology. Further rapid advances are anticipated. Chem. 1992, 338, 85 – 98; d) M. Rudolph, J. Electroanal. Chem. 1994,
375, 89 – 99; e) M. Rudolph, J. Electroanal. Chem. 2001, 503, 15 – 27;
f) M. Rudolph, J. Electroanal. Chem. 2002, 529, 97 – 108; g) M. Rudolph,
J. Electroanal. Chem. 2003, 543, 23 – 39; h) M. Rudolph, J. Electroanal.
Acknowledgements Chem. 2004, 571, 289 – 307.
[2] Digisim, in http://www.basinc.com/products/ec/digisim/.
[3] a) R. G. Compton, E. Laborda, K. R. Ward, Understanding Voltammetry:
The research leading to these results has received partial funding Simulation of Electrode Processes, Imperial College Press, London,
from the European Research Council (ERC) under the European 2014; b) D. Britz, Digital Simulation in Electrochemistry, Springer, Berlin,
Union’s Seventh Framework Programme (FP/2007-2013), ERC 2005; c) E. J. Dickinson, H. Ekstrçm, E. Fontes, Electrochem. Commun.
Grant Agreement no. 320403. E. O. B. thanks the Engineering and 2014, 40, 71 – 74.
[4] a) A. M. Bond, E. A. Mashkina, A. N. Simonov, Developments in Electro-
Physical Sciences Research Council (EPSRC) for a ‘Doctoral Prize’.
chemistry, John Wiley & Sons, Ltd, Chichester, 2014, pp. 21 – 47; b) L. K.
A. M. and E. L. greatly appreciate the financial support provided Bieniasz, Comput. Chem. 1992, 16, 11 – 14; c) L. K. Bieniasz, Comput.
by the Ministerio de Economia y Competitividad, Spain (Project Chem. 1997, 21, 1 – 12.
Number CTQ2012-36700, cofunded by the European Regional De- [5] S. E. Creager, T. T. Wooster, Anal. Chem. 1998, 70, 4257 – 4263.
velopment Fund) and by the Fundaciûn S¦neca de la Regiûn de [6] R. G. Compton, A. C. Fisher, R. Geoffrey Wellington, P. J. Dobson, P. A.
Leigh, J. Phys. Chem. 1993, 97, 10410 – 10415.
Murcia (Project 18968/JLI/13). E. L. also thanks the funding re- [7] J. M. Sav¦ant, Elements of Molecular and Biomolecular Electrochemistry:
ceived from the European Union Seventh Framework Pro- An Electrochemical Approach to Electron Transfer Chemistry, Wiley,
gramme–Marie Curie COFUND (FP7/2007-2013) under University Hoboken, NJ, 2006.
of Murcia Incoming Mobility Programme Action (U-IMPACT). [8] a) D. H. Evans, Chem. Rev. 1990, 90, 739 – 751; b) D. H. Evans, Chem.
Rev. 2008, 108, 2113 – 2144.
[9] a) R. Guidelli, R. G. Compton, J. M. Feliu, E. Gileadi, J. Lipkowski, W.
Keywords: cyclic voltammetry · electrochemistry · electron Schmickler, S. Trasatti, Pure Appl. Chem. 2014, 86, 245 – 258; b) R.
transfer · kinetics · mass transport · nanoparticle Guidelli, R. G. Compton, J. M. Feliu, E. Gileadi, J. Lipkowski, W. Schmick-
ler, S. Trasatti, Pure Appl. Chem. 2014, 86, 259 – 262.
electrochemistry · stripping voltammetry

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 256 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[10] O. V. Klymenko, R. G. Compton, J. Electroanal. Chem. 2004, 571, 207 – F. L. Bhatti, T. J. Donohoe, R. G. Compton, Electroanalysis 2007, 19, 11 –
210. 22; c) L. Xiong, L. Aldous, M. C. Henstridge, R. G. Compton, Anal. Meth-
[11] M. C. Henstridge, R. G. Compton, J. Electroanal. Chem. 2012, 681, 109 – ods 2012, 4, 371 – 376.
112. [43] a) L. E. Barrosse-Antle, L. Aldous, C. Hardacre, A. M. Bond, R. G. Comp-
[12] C. Batchelor-McAuley, R. G. Compton, J. Phys. Chem. C 2014, 118, ton, J. Phys. Chem. C 2009, 113, 7750 – 7754; b) L. E. Barrosse-Antle, C.
30034 – 30038. Hardacre, R. G. Compton, J. Phys. Chem. B 2009, 113, 2805 – 2809;
[13] A. C. Testa, W. H. Reinmuth, Anal. Chem. 1961, 33, 1320 – 1324. c) L. E. Barrosse-Antle, C. Hardacre, R. G. Compton, J. Phys. Chem. B
[14] a) I. Noviandri, K. N. Brown, D. S. Fleming, P. T. Gulyas, P. A. Lay, A. F. 2009, 113, 1007 – 1011; d) Y. Meng, L. Aldous, R. G. Compton, J. Phys.
Masters, L. Phillips, J. Phys. Chem. B 1999, 103, 6713 – 6722; b) F. Bar- Chem. C 2011, 115, 14334 – 14340.
riÀre, W. E. Geiger, J. Am. Chem. Soc. 2006, 128, 3980 – 3989. [44] R. G. Evans, O. V. Klymenko, P. D. Price, S. G. Davies, C. Hardacre, R. G.
[15] M. Quan, D. Sanchez, M. F. Wasylkiw, D. K. Smith, J. Am. Chem. Soc. Compton, ChemPhysChem 2005, 6, 526 – 533.
2007, 129, 12847 – 12856. [45] M. C. Buzzeo, O. V. Klymenko, J. D. Wadhawan, C. Hardacre, K. R.
[16] Q. Li, C. Batchelor-McAuley, N. S. Lawrence, R. S. Hartshorne, R. G. Seddon, R. G. Compton, J. Phys. Chem. A 2003, 107, 8872 – 8878.
Compton, Chem. Commun. 2011, 47, 11426 – 11428. [46] U. Schrçder, J. D. Wadhawan, R. G. Compton, F. Marken, P. A. Z. Suarez,
[17] C. Batchelor-McAuley, Q. Li, S. M. Dapin, R. G. Compton, J. Phys. Chem. C. S. Consorti, R. F. De Souza, J. Dupont, New J. Chem. 2000, 24, 1009 –
B 2010, 114, 4094 – 4100. 1015.
[18] E. Laborda, A. Molina, Q. Li, C. Batchelor-McAuley, R. G. Compton, Phys. [47] R. G. Evans, O. V. Klymenko, C. Hardacre, K. R. Seddon, R. G. Compton,
Chem. Chem. Phys. 2012, 14, 8319 – 8327. J. Electroanal. Chem. 2003, 556, 179 – 188.
[19] N. A. Mac†as-Ruvalcaba, D. H. Evans, J. Phys. Chem. B 2005, 109, [48] G. D. Allen, M. C. Buzzeo, C. Villagr‚n, C. Hardacre, R. G. Compton, J.
14642 – 14647. Electroanal. Chem. 2005, 575, 311 – 320.
[20] D. H. Evans, M. W. Lehmann, Acta Chem. Scand. 1999, 53, 765 – 774. [49] a) D. S. Silvester, A. J. Wain, L. Aldous, C. Hardacre, R. G. Compton, J.
[21] C. Batchelor-McAuley, R. G. Compton, J. Electroanal. Chem. 2012, 669, Electroanal. Chem. 2006, 596, 131 – 140; b) S. Ernst, K. R. Ward, S. E.
73 – 81. Norman, C. Hardacre, R. G. Compton, Phys. Chem. Chem. Phys. 2013,
[22] I. Streeter, S. F. Jenkinson, G. W. J. Fleet, R. G. Compton, J. Electroanal. 15, 6382 – 6389.
Chem. 2007, 600, 285 – 293. [50] R. G. Evans, R. G. Compton, ChemPhysChem 2006, 7, 488 – 496.
[23] C. Costentin, M. Robert, J. M. Sav¦ant, Chem. Rev. 2010, 110, PR1 – [51] T. L. Broder, D. S. Silvester, L. Aldous, C. Hardacre, R. G. Compton, J.
PR40. Phys. Chem. B 2007, 111, 7778 – 7785.
[24] Q. Li, C. Batchelor-McAuley, R. G. Compton, J. Phys. Chem. B 2010, 114,
[52] E. I. Rogers, I. Streeter, L. Aldous, C. Hardacre, R. G. Compton, J. Phys.
7423 – 7428.
Chem. C 2008, 112, 10976 – 10981.
[25] a) E. Laviron, J. Electroanal. Chem. Interfacial Electrochem. 1979, 101,
[53] J. S. Long, D. S. Silvester, A. S. Barnes, N. V. Rees, L. Aldous, C. Hardacre,
19 – 28; b) Y. Xiao, F. Patolsky, E. Katz, J. F. Hainfeld, I. Willner, Science
R. G. Compton, J. Phys. Chem. C 2008, 112, 6993 – 7000.
2003, 299, 1877 – 1881.
[54] S. R. Belding, N. V. Rees, L. Aldous, C. Hardacre, R. G. Compton, J. Phys.
[26] R. G. Compton, C. E. Banks, Understanding Voltammetry, 2nd ed., Impe-
Chem. C 2008, 112, 1650 – 1657.
rial College Press, London, 2011.
[55] R. Wibowo, S. E. W. Jones, R. G. Compton, J. Chem. Eng. Data 2010, 55,
[27] J. A. Alden, R. G. Compton, Anal. Chem. 2000, 72, 198A – 203A.
1374 – 1376.
[28] C. C. M. Neumann, E. Laborda, K. Tschulik, K. R. Ward, R. G. Compton,
[56] Y. Wang, E. I. Rogers, S. R. Belding, R. G. Compton, J. Electroanal. Chem.
Nano Res. 2013, 6, 511 – 524.
2010, 648, 134 – 142.
[29] a) C. Batchelor-McAuley, K. Tschulik, R. G. Compton, Port. Electrochim.
[57] Y. Wang, S. R. Belding, E. I. Rogers, R. G. Compton, J. Electroanal. Chem.
Acta 2013, 31, 249 – 256; b) C. Batchelor-McAuley, K. Tschulik, C. C. M.
2011, 650, 196 – 204.
Neumann, E. Laborda, R. G. Compton, Int. J. Electrochem. Sci. 2014, 9,
[58] a) Y. Meng, L. Aldous, S. R. Belding, R. G. Compton, Chem. Commun.
1132 – 1138.
2012, 48, 5572 – 5574; b) Y. Meng, L. Aldous, S. R. Belding, R. G. Comp-
[30] M. C. Buzzeo, R. G. Evans, R. G. Compton, ChemPhysChem 2004, 5,
ton, Phys. Chem. Chem. Phys. 2012, 14, 5222 – 5228.
1106 – 1120.
[59] I. Streeter, M. Thompson, R. G. Compton, J. Electroanal. Chem. 2006,
[31] a) G. Gebresilassie Eshetu, M. Armand, B. Scrosati, S. Passerini, Angew.
Chem. Int. Ed. 2014, 53, 13342 – 13359; Angew. Chem. 2014, 126, 591, 133 – 140.
13558 – 13576; b) D. Bresser, E. Paillard, S. Passerini, J. Electrochem. Sci. [60] E. E. L. Tanner, L. Xiong, E. O. Barnes, R. G. Compton, J. Electroanal.
Tech. 2014, 5, 37 – 44. Chem. 2014, 727, 59 – 68.
[32] L. Xiong, R. G. Compton, Int. J. Electrochem. Sci. 2014, 9, 7152 – 7181. [61] a) E. Laborda, M. C. Henstridge, C. Batchelor-McAuley, R. G. Compton,
[33] A. A. Kornyshev, L. P. Yang, Electrochem. Commun. 2014, 48, 173 – 177. Chem. Soc. Rev. 2013, 42, 4894 – 4905; b) L. Fan, Y. Liu, J. Xiong, H. S.
[34] M. V. Fedorov, A. A. Kornyshev, Chem. Rev. 2014, 114, 2978 – 3036. White, S. Chen, ACS Nano 2014, 8, 10426 – 10436; c) S. Amemiya, N.
[35] a) D. S. Silvester, E. I. Rogers, L. E. Barrosse-Antle, T. L. Broder, R. G. Nioradze, P. Santhosh, M. J. Deible, Anal. Chem. 2011, 83, 5928 – 5935;
Compton, J. Braz. Chem. Soc. 2008, 19, 611 – 620; b) D. S. Silvester, Ana- d) S. Chen, Y. Liu, J. Chen, Chem. Soc. Rev. 2014, 43, 5372 – 5386;
lyst 2011, 136, 4871 – 4882; c) L. E. Barrosse-Antle, A. M. Bond, R. G. e) S. W. Feldberg, Anal. Chem. 2010, 82, 5176 – 5183; f) L.-B. Zhao, A. K.
Compton, A. M. O’Mahony, E. I. Rogers, D. S. Silvester, Chem. Asian J. Mishra, D. H. Waldeck, J. Phys. Chem. C 2013, 117, 20746 – 20761.
2010, 5, 202 – 230. [62] a) J. A. V. Butler, Trans. Faraday Soc. 1924, 19, 729 – 733; b) T. Erdey-
[36] a) E. I. Rogers, B. S̃ljukic, C. Hardacre, R. G. Compton, J. Chem. Eng. Grfflz, M. Volmer, Z. Physik. Chem. 1930, 150A, 203 – 213.
Data 2009, 54, 2049 – 2053; b) A. M. O’Mahony, D. S. Silvester, L. [63] a) R. A. Marcus, J. Chem. Phys. 1956, 24, 966 – 978; b) N. S. Hush, J.
Aldous, C. Hardacre, R. G. Compton, J. Chem. Eng. Data 2008, 53, Chem. Phys. 1958, 28, 962 – 972.
2884 – 2891. [64] a) J.-M. Sav¦ant, D. Tessier, Faraday Discuss. Chem. Soc. 1982, 74, 57 –
[37] M. C. Buzzeo, C. Hardacre, R. G. Compton, ChemPhysChem 2006, 7, 72; b) C. E. D. Chidsey, Science 1991, 251, 919 – 922; c) R. M. Haddox,
176 – 180. H. O. Finklea, J. Phys. Chem. B 2004, 108, 1694 – 1700; d) S. Fletcher,
[38] X. J. Huang, E. I. Rogers, C. Hardacre, R. G. Compton, J. Phys. Chem. B T. S. Varley, Phys. Chem. Chem. Phys. 2011, 13, 5359 – 5364.
2009, 113, 8953 – 8959. [65] a) E. Laborda, M. C. Henstridge, R. G. Compton, J. Electroanal. Chem.
[39] A. M. O’Mahony, D. S. Silvester, L. Aldous, C. Hardacre, R. G. Compton, 2012, 667, 48 – 53; b) M. C. Henstridge, E. Laborda, R. G. Compton, J.
J. Phys. Chem. C 2008, 112, 7725 – 7730. Electroanal. Chem. 2012, 674, 90 – 96; c) M. C. Henstridge, E. Laborda, Y.
[40] E. I. Rogers, D. S. Silvester, D. L. Poole, L. Aldous, C. Hardacre, R. G. Wang, D. Suwatchara, N. Rees, Ý. Molina, F. Mart†nez-Ortiz, R. G. Comp-
Compton, J. Phys. Chem. C 2008, 112, 2729 – 2735. ton, J. Electroanal. Chem. 2012, 672, 45 – 52.
[41] A. S. Barnes, E. I. Rogers, I. Streeter, L. Aldous, C. Hardacre, R. G. Comp- [66] R. A. Marcus, J. Chem. Phys. 1965, 43, 679 – 701.
ton, J. Phys. Chem. B 2008, 112, 7560 – 7565. [67] W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical
[42] a) O. V. Klymenko, R. G. Evans, C. Hardacre, I. B. Svir, R. G. Compton, J. Recipes 3rd edition: The Art of Scientific Computing, 3rd ed., Cambridge
Electroanal. Chem. 2004, 571, 211 – 221; b) C. A. Paddon, D. S. Silvester, University Press, New York, 2007.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 257 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[68] a) K. B. Oldham, J. C. Myland, J. Electroanal. Chem. 2011, 655, 65 – 72; chim. Acta 2011, 56, 7189 – 7193; i) V. Mirceski, R. Gulaboski, M. Lovrić,
b) A. Migliore, A. Nitzan, J. Electroanal. Chem. 2012, 671, 99 – 101; c) Y. I. Bogeski, R. Kappl, M. Hoth, Electroanalysis 2013, 25, 2411 – 2422; j) R.
Zeng, R. B. Smith, P. Bai, M. Z. Bazant, J. Electroanal. Chem. 2014, 735, Gulaboski, V. Mirčeski, I. Bogeski, M. Hoth, J. Solid State Electrochem.
77 – 83. 2012, 16, 2315 – 2328; k) R. Gulaboski, L. Mihajlov, Biophys. Chem.
[69] N. Fatouros, D. Krulic, J. Electroanal. Chem. 2002, 520, 1 – 5. 2011, 155, 1 – 9; l) R. Gulaboski, J. Solid State Electrochem. 2009, 13,
[70] a) E. Laborda, M. C. Henstridge, A. Molina, F. Mart†nez-Ortiz, R. G. 1015 – 1024; m) V. Mirceski, S. B. Hocevar, B. Ogorevc, R. Gulaboski, I.
Compton, J. Electroanal. Chem. 2011, 660, 169 – 177; b) G. P. Stevenson, Drangov, Anal. Chem. 2012, 84, 4429 – 4436; n) V. Mirceski, B. Sebez, M.
R. E. Baker, G. F. Kennedy, A. M. Bond, D. J. Gavaghan, K. Gillow, Phys. Jancovska, B. Ogorevc, S. B. Hocevar, Electrochim. Acta 2013, 105, 254 –
Chem. Chem. Phys. 2013, 15, 2210 – 2221. 260; o) V. Mirčeski, A. Bobrowski, J. Zarebski, F. Spasovski, Electrochim.
[71] E. Laborda, Y. Wang, M. C. Henstridge, F. Mart†nez-Ortiz, A. Molina, R. G. Acta 2010, 55, 8696 – 8703.
Compton, Chem. Phys. Lett. 2011, 512, 133 – 137. [95] M. A. Mann, J. C. Helfrick, L. A. Bottomley, Anal. Chem. 2014, 86, 8183 –
[72] M. C. Henstridge, E. Laborda, E. J. F. Dickinson, R. G. Compton, J. Elec- 8191.
troanal. Chem. 2012, 664, 73 – 79. [96] V. Mirčeski, M. Lovrić, J. Electroanal. Chem. 2001, 497, 114 – 124.
[73] M. C. Henstridge, K. R. Ward, R. G. Compton, J. Electroanal. Chem. 2014, [97] D. Jadreško, M. Zelić, M. Lovrić, J. Electroanal. Chem. 2010, 645, 103 –
712, 14 – 18. 108.
[74] M. C. Henstridge, N. V. Rees, R. G. Compton, J. Electroanal. Chem. 2012, [98] V. Mirceski, E. Laborda, D. Guziejewski, R. G. Compton, Anal. Chem.
687, 79 – 83. 2013, 85, 5586 – 5594.
[75] E. Laborda, M. C. Henstridge, R. G. Compton, J. Electroanal. Chem. [99] a) M. Zelić, Croat. Chem. Acta 2006, 79, 49 – 55; b) M. Zelić, M. Lovrić, J.
2012, 681, 96 – 102. Electroanal. Chem. 2009, 637, 28 – 32; c) M. Lovrić, D. Jadreško, Electro-
[76] D. V. Matyushov, J. Chem. Phys. 2009, 130, 234704. chim. Acta 2010, 55, 948 – 951.
[77] a) E. Laborda, E. I. Rogers, F. Mart†nez-Ortiz, Ý. Molina, R. G. Compton, [100] D. Jadreško, M. Zelić, J. Electroanal. Chem. 2013, 707, 20 – 30.
Electroanalysis 2010, 22, 2784 – 2793; b) E. Laborda, E. I. Rogers, F. Mar- [101] a) G. C. Barker, Anal. Chim. Acta 1958, 18, 118 – 131; b) J. H. Christie, P. J.
t†nez-Ortiz, Ý. Molina, R. G. Compton, Electrochim. Acta 2010, 55, Lingane, J. Electroanal. Chem. 1965, 10, 176 – 182.
6577 – 6585; c) Ý. Molina, F. Mart†nez-Ortiz, E. Laborda, R. G. Compton, [102] a) R. Bilewicz, R. A. Osteryoung, J. Osteryoung, Anal. Chem. 1986, 58,
Electrochim. Acta 2010, 55, 5163 – 5172; d) Ý. Molina, F. Mart†nez-Ortiz, 2761 – 2765; b) R. Bilewicz, K. Wikiel, R. Osteryoung, J. Osteryoung,
E. Laborda, R. G. Compton, J. Electroanal. Chem. 2010, 648, 67 – 77. Anal. Chem. 1989, 61, 965 – 972; c) M. Seralathan, R. Osteryoung, J. Os-
[78] A. J. Bard, L. R. Faulkner, Electrochemical Methods: Fundamentals and teryoung, J. Electroanal. Chem. 1986, 214, 141 – 156; d) M. Seralathan,
Applications, 2nd ed., John Wiley & Sons Inc., Hoboken, NJ, 2001. R. A. Osteryoung, J. G. Osteryoung, J. Electroanal. Chem. 1987, 222,
[79] a) E. Laborda, F. Mart†nez-Ortiz, Ý. Molina, Electroanalysis 2010, 22, 69 – 100.
1857 – 1866; b) E. Laborda, F. Mart†nez-Ortiz, Ý. Molina, Electrochim. [103] U. Kalapathy, D. E. Tallman, Anal. Chem. 1992, 64, 2693 – 2700.
Acta 2011, 56, 5335 – 5342. [104] E. J. F. Dickinson, J. G. Limon-Petersen, N. V. Rees, R. G. Compton, J.
[80] A. Molina, E. Laborda, F. Martinez-Ortiz, E. Torralba, R. Compton, Elec- Phys. Chem. C 2009, 113, 11157 – 11171.
trochim. Acta Electrochimica Acta 2013, 87, 416 – 424. [105] A. S. Barnes, I. Streeter, R. G. Compton, J. Electroanal. Chem. 2008, 623,
[81] A. Molina, J. Gonz‚lez, E. Laborda, Y. Wang, R. G. Compton, Phys. 129 – 133.
Chem. Chem. Phys. 2011, 13, 14694 – 14704. [106] a) M. Lûpez-Ten¦s, C. Serna, M. M. Moreno, Ý. Molina, J. Electroanal.
[82] a) E. Laborda, J. Olmos, A. Molina, J. Electroanal. Chem. 2015, 741, Chem. 2007, 603, 249 – 259.
140 – 148; b) E. Laborda, J. M. Olmos, E. Torralba, A. Molina, Anal. [107] H. A. Heering, M. S. Mondal, F. A. Armstrong, Anal. Chem. 1999, 71,
Chem. 2015, 87, 1676. 174 – 182.
[83] a) A. Molina, J. Gonzalez, E. Laborda, R. Compton, Int. J. Electrochem. [108] B. Hai, Y. V. Tolmachev, K. A. Loparo, C. Zanelli, D. Scherson, J. Electro-
Sci. 2012, 7, 5765 – 5778; b) A. Molina, J. Gonzalez, E. Laborda, Q. Li, C. chem. Soc. 2011, 158, F15 – F19.
Batchelor-McAuley, R. G. Compton, J. Phys. Chem. C 2012, 116, 1070 – [109] Q. Li, C. Batchelor-McAuley, N. S. Lawrence, R. S. Hartshorne, R. G.
1079. Compton, New J. Chem. 2011, 35, 2462 – 2470.
[84] A. Molina, M. M. Moreno, C. Serna, L. Camacho, Electrochem. Commun. [110] a) B. R. Kozub, M. C. Henstridge, C. Batchelor-McAuley, R. G. Compton,
2001, 3, 324 – 329. Int. J. Electrochem. Sci. 2011, 6, 6047 – 6062; b) B. R. Kozub, M. C. Hen-
[85] W. S. Go, J. J. O’Dea, J. Osteryoung, J. Electroanal. Chem. 1988, 255, stridge, C. Batchelor-McAuley, R. G. Compton, ChemPhysChem 2011,
21 – 44. 12, 2806 – 2815.
[86] a) A. Molina, C. Serna, Q. Li, E. Laborda, C. Batchelor-McAuley, R. G. [111] C. Amatore, S. Szunerits, L. Thouin, J.-S. Warkocz, J. Electroanal. Chem.
Compton, J. Phys. Chem. C 2012, 116, 11470 – 11479; b) A. Molina, J. 2001, 500, 62 – 70.
Gonzalez, E. Laborda, R. G. Compton, Russ. J. Electrochem. 2012, 48, [112] S. Chen, Y. Liu, Phys. Chem. Chem. Phys. 2014, 16, 635 – 652.
600 – 609. [113] M. C. Henstridge, R. G. Compton, Chem. Rec. 2012, 12, 63 – 71.
[87] Ý. Molina, E. Laborda, E. I. Rogers, F. Mart†nez-Ortiz, C. Serna, J. G. [114] a) A. Molina, J. Gonz‚lez, F. Mart†nez-Ortiz, R. G. Compton, J. Phys.
Limon-Petersen, N. V. Rees, R. G. Compton, J. Electroanal. Chem. 2009, Chem. C 2010, 114, 4093 – 4099; b) A. Molina, E. Laborda, J. Gonz‚lez,
634, 73 – 81. R. G. Compton, Phys. Chem. Chem. Phys. 2013, 15, 7106 – 7113; c) A.
[88] R. L. Birke, M.-H. Kim, M. Strassfeld, Anal. Chem. 1981, 53, 852 – 856. Molina, J. Gonzalez, M. C. Henstridge, R. G. Compton, Electrochim. Acta
[89] M. Zelić, M. Lovrić, Electroanalysis 2011, 23, 642 – 650. 2011, 56, 4589 – 4594; d) A. Molina, J. Gonz‚lez, E. Laborda, R. G.
[90] E. Laborda, J. Gonz‚lez, Ý. Molina, Electrochem. Commun. 2014, 43, Compton, Phys. Chem. Chem. Phys. 2013, 15, 2381 – 2388.
25 – 30. [115] A. Fick, Annalen der Physik 1855, 170, 59 – 86.
[91] A. Molina, E. Laborda, F. Martinez-Ortiz, D. F. Bradley, D. J. Schiffrin, [116] T. Fließbach, Statistische Physik, Lehrbuch zur theoretischen Physik IV,
R. G. Compton, J. Electroanal. Chem. 2011, 659, 12 – 24. 3. Auflage, Spektrum Akademischer Verlag, Heidelberg, 1999.
[92] J. Osteryoung, E. Kirowa-Eisner, Anal. Chem. 1980, 52, 62 – 66. [117] E. K•telhçn, K. J. Krause, P. S. Singh, S. G. Lemay, B. Wolfrum, J. Am.
[93] Ý. Molina, R. G. Compton, C. Serna, F. Mart†nez-Ortiz, E. Laborda, Elec- Chem. Soc. 2013, 135, 8874 – 8881.
trochim. Acta 2009, 54, 2320 – 2328. [118] a) C. Amatore, F. Grìn, E. Maisonhaute, Angew. Chem. Int. Ed. 2003, 42,
[94] a) V. Mirceski, S. Komorsky-Lovrić, M. Lovrić, Square-Wave Voltammetry: 4944 – 4947; Angew. Chem. 2003, 115, 5094 – 5097; b) S. Liang, X. Dong,
Theory and Application, 2nd ed., Springer, Berlin, 2007; b) Š. Komorsky- J. Solid State Electrochem. 2008, 12, 701 – 706.
Lovrić, M. Lovrić, J. Electroanal. Chem. 2011, 660, 22 – 25; c) Š. Komor- [119] I. J. Cutress, R. G. Compton, Chem. Phys. Lett. 2011, 508, 306 – 313.
sky-Lovrić, M. Lovrić, Electrochim. Acta 2012, 69, 60 – 64; d) S. Komor- [120] I. J. Cutress, E. J. Dickinson, R. G. Compton, J. Electroanal. Chem. 2011,
sky-Lovrić, M. Lovrić, Int. J. Electrochem. Sci. 2014, 9, 435 – 444; e) M. 655, 1 – 8.
Lovrić, S. Komorsky-Lovrić, Int. J. Electrochem. Sci. 2014, 9, 5549 – 5559; [121] E. K•telhçn, K. J. Krause, B. Wolfrum, R. G. Compton, ChemPhysChem
f) Š. Komorsky-Lovrić, D. Jadreško, M. Lovrić, Electrochim. Acta 2014, 2014, 15, 872 – 875.
130, 286 – 289; g) M. Lovrić, D. Jadreško, Š. Komorsky-Lovrić, Electro- [122] P. S. Singh, E. K•telhçn, K. Mathwig, B. Wolfrum, S. G. Lemay, ACS Nano
chim. Acta 2013, 90, 226 – 231; h) Š. Komorsky-Lovrić, M. Lovrić, Electro- 2012, 6, 9662 – 9671.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 258 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[123] E. K•telhçn, K. J. Krause, K. Mathwig, S. G. Lemay, B. Wolfrum, ACS [151] a) T. J. Davies, R. G. Compton, J. Electroanal. Chem. 2005, 585, 63 – 82;
Nano 2014, 8, 4924 – 4930. b) I. Streeter, R. Baron, R. G. Compton, J. Phys. Chem. C 2007, 111,
[124] M. A. Zevenbergen, P. S. Singh, E. D. Goluch, B. L. Wolfrum, S. G. Lemay, 17008 – 17014; c) S. R. Belding, R. G. Compton, J. Phys. Chem. C 2010,
Anal. Chem. 2009, 81, 8203 – 8212. 114, 8309 – 8319.
[125] a) S. J. Kwon, H. Zhou, F.-R. F. Fan, V. Vorobyev, B. Zhang, A. J. Bard, [152] a) P. A. Bobbert, M. M. Wind, J. Vlieger, Physica 1987, 141, 58 – 72; b) I.
Phys. Chem. Chem. Phys. 2011, 13, 5394 – 5402; b) E. K•telhçn, R. G. Streeter, R. G. Compton, J. Phys. Chem. C 2007, 111, 18049 – 18054.
Compton, Chem. Sci. 2014, 5, 4592 – 4598. [153] N. Godino, X. Borris¦, F. X. MuÇoz, F. J. del Campo, R. G. Compton, J.
[126] M. A. Bevan, D. C. Prieve, J. Chem. Phys. 2000, 113, 1228 – 1236. Phys. Chem. C 2009, 113, 11119 – 11125.
[127] A. Goldman, R. G. Cox, H. Brenner, Chem. Eng. Sci. 1967, 22, 637 – 651. [154] J. L. Fern‚ndez, M. Wijesinghe, C. G. Zoski, Anal. Chem. 2015, 87,
[128] P. H. Rieger, Electrochemistry, Prentice-Hall, Inc., Englewood Cliffs, NJ, 1066 – 1074.
1987. [155] Y. G. Zhou, F. W. Campbell, S. R. Belding, R. G. Compton, Chem. Phys.
[129] a) E. O. Barnes, S. R. Belding, R. G. Compton, J. Electroanal. Chem. 2011, Lett. 2010, 497, 200 – 204.
660, 185 – 194; b) E. O. Barnes, Y. Wang, S. R. Belding, R. G. Compton, [156] M. Gara, E. Laborda, P. Holdway, A. Crossley, C. J. V. Jones, R. G. Comp-
ChemPhysChem 2012, 13, 92 – 95; c) E. O. Barnes, Y. Wang, J. G. Limon- ton, Phys. Chem. Chem. Phys. 2013, 15, 19487 – 19495.
Petersen, S. R. Belding, R. G. Compton, J. Electroanal. Chem. 2011, 659, [157] Y. Wang, K. R. Ward, E. Laborda, C. Salter, A. Crossley, R. M. J. Jacobs,
25 – 35; d) S. R. Belding, J. G. Limon-Petersen, E. J. F. Dickinson, R. G. R. G. Compton, Small 2013, 9, 478 – 486.
Compton, Angew. Chem. Int. Ed. 2010, 49, 9242 – 9245; Angew. Chem. [158] Y. Wang, E. Laborda, K. Tschulik, C. Damm, A. Molina, R. G. Compton,
2010, 122, 9428 – 9431; e) J. G. Limon-Petersen, I. Streeter, N. V. Rees, Nanoscale 2014, 6, 11024 – 11030.
R. G. Compton, J. Phys. Chem. C 2008, 112, 17175 – 17182; f) Y. Wang, [159] a) O. S. Ivanova, F. P. Zamborini, J. Am. Chem. Soc. 2010, 132, 70 – 72;
E. O. Barnes, R. G. Compton, ChemPhysChem 2012, 13, 3441 – 3444; b) K. Z. Brainina, L. G. Galperin, T. Y. Kiryuhina, A. L. Galperin, N. Y.
g) Y. Wang, E. O. Barnes, E. Laborda, A. Molina, R. G. Compton, J. Elec- Stozhko, A. M. Murzakaev, O. R. Timoshenkova, J. Solid State Electro-
troanal. Chem. 2012, 673, 13 – 23. chem. 2012, 16, 2365 – 2372.
[130] E. J. F. Dickinson, J. G. Limon-Petersen, R. G. Compton, J. Solid State [160] a) R. A. Masitas, I. V. Khachian, B. L. Bill, F. P. Zamborini, Langmuir 2014,
Electrochem. 2011, 15, 1335 – 1345. 30, 13075 – 13084; b) K. Z. Brainina, L. G. Galperin, E. V. Vikulova, J. Solid
[131] a) E. J. F. Dickinson, R. G. Compton, Chem. Phys. Lett. 2010, 497, 178 – State Electrochem. 2012, 16, 2357 – 2363.
183; b) I. Streeter, R. G. Compton, J. Phys. Chem. C 2008, 112, 13716 – [161] H. S. Toh, K. Jurkschat, R. G. Compton, Chem. Eur. J. 2015, 21, 2998 –
13728. 3004.
[132] a) J. G. Limon-Petersen, E. J. F. Dickinson, S. R. Belding, N. V. Rees, R. G. [162] a) S. E. Ward Jones, F. W. Campbell, R. Baron, L. Xiao, R. G. Compton, J.
Compton, J. Electroanal. Chem. 2010, 650, 135 – 142; b) J. G. Limon-Pe- Phys. Chem. C 2008, 112, 17820 – 17827; b) S. E. Ward Jones, S. H.
tersen, J. T. Han, N. V. Rees, E. J. F. Dickinson, I. Streeter, R. G. Compton, Zheng, C. A. Jeffrey, S. Seretis, S. Morin, R. G. Compton, J. Electroanal.
J. Phys. Chem. C 2010, 114, 2227 – 2236; c) J. G. Limon-Petersen, I. Chem. 2008, 616, 38 – 44.
Streeter, N. V. Rees, R. G. Compton, J. Phys. Chem. C 2009, 113, 333 – [163] H. S. Toh, C. Batchelor-McAuley, K. Tschulik, M. Uhlemann, A. Crossley,
337. R. G. Compton, Nanoscale 2013, 5, 4884 – 4893.
[133] V. Levich, Dokl. Akad. Nauk SSSR 1949, 67, 309 – 312. [164] C. C. M. Neumann, C. Batchelor-McAuley, K. Tschulik, H. S. Toh, P.
[134] A. Frumkin, Z. Physik. Chem. 1933, 164A, 121 – 133. Shumbula, J. Pillay, R. Tshikhudo, R. G. Compton, ChemElectroChem.
[135] E. J. F. Dickinson, R. G. Compton, J. Electroanal. Chem. 2011, 661, 198 – 2014, 1, 87 – 89.
212. [165] C. Batchelor-McAuley, A. Martinez-Marrades, K. Tschulik, A. N. Patel, C.
[136] K. B. Oldham, C. G. Zoski, Mass Transport to Electrodes In Comprehensive Combellas, F. Kanoufi, G. Tessier, R. G. Compton, Chem. Phys. Lett.
Chemical Kinetics, Vol. 26, Elsevier, Amsterdam, 1986. 2014, 597, 20 – 25.
[137] S. R. Belding, R. G. Compton, J. Electroanal. Chem. 2012, 683, 1 – 13. [166] a) J. C. Harfield, K. E. Toghill, C. Batchelor-McAuley, C. Downing, R. G.
[138] C. P. Andrieux, J. M. Sav¦ant, J. Electroanal. Chem. 1970, 28, 339 – 348. Compton, Electroanalysis 2011, 23, 931 – 938; b) L. Xiao, G. G. Wild-
[139] a) C. A. Amatore, A. Jutand, F. Pfluger, J. Electroanal. Chem. 1987, 218, goose, R. G. Compton, Anal. Chim. Acta 2008, 620, 44 – 49.
361 – 365. [167] a) H. S. Toh, C. Batchelor-McAuley, K. Tschulik, R. G. Compton, Analyst
[140] a) R. G. Compton, R. A. W. Dryfe, J. Electroanal. Chem. 1994, 375, 247 – 2013, 138, 4292 – 4297; b) H. S. Toh, K. Tschulik, C. Batchelor-McAuley,
255; b) J. G. Lawless, M. D. Hawley, J. Electroanal. Chem. 1969, 21, 365 – R. G. Compton, Analyst 2014, 139, 3986 – 3990.
375; c) R. F. Nelson, A. K. Carpenter, E. T. Seo, J. Electrochem. Soc. 1973, [168] a) U. Hasse, S. Fletcher, F. Scholz, J. Solid State Electrochem. 2006, 10,
120, 206 – 210. 833 – 840; b) U. Hasse, F. Scholz, Electrochem. Commun. 2004, 6, 409 –
[141] R. Mìller, L. Lamberts, M. Evers, J. Electroanal. Chem. 1996, 417, 35 – 43. 412.
[142] C. Amatore, J. M. Sav¦ant, J. Electroanal. Chem. 1978, 86, 227 – 232. [169] A. B. Garrett, M. V. Noble, S. Miller, J. Chem. Educ. 1942, 19, 485.
[143] B. R. Kozub, N. V. Rees, R. G. Compton, Sens. Actuators B 2010, 143, [170] N. V. Rees, Electrochem. Commun. 2014, 43, 83 – 86.
539 – 546. [171] a) M. Pumera, ACS Nano 2014, 8, 7555 – 7558; b) S. E. F. Kleijn, S. C. S.
[144] a) I. Streeter, G. G. Wildgoose, L. Shao, R. G. Compton, Sens. Actuators B Lai, M. T. M. Koper, P. R. Unwin, Angew. Chem. Int. Ed. 2014, 53, 3558 –
2008, 133, 462 – 466; b) M. C. Henstridge, E. J. F. Dickinson, M. Aslano- 3586; Angew. Chem. 2014, 126, 3630 – 3660.
glu, C. Batchelor-McAuley, R. G. Compton, Sens. Actuators B 2010, 145, [172] a) R. Dasari, K. Tai, D. A. Robinson, K. J. Stevenson, ACS Nano 2014, 8,
417 – 427; c) M. J. Sims, N. V. Rees, E. J. F. Dickinson, R. G. Compton, 4539 – 4546; b) R. Dasari, B. Walther, D. A. Robinson, K. J. Stevenson,
Sens. Actuators B 2010, 144, 153 – 158; d) M. C. Henstridge, E. J. F. Dick- Langmuir 2013, 29, 15100 – 15106; c) T. M. Alligrant, M. J. Anderson, R.
inson, R. G. Compton, Russ. J. Electrochem. 2012, 48, 629 – 635. Dasari, K. J. Stevenson, R. M. Crooks, Langmuir 2014, 30, 13462 – 13469.
[145] a) K. R. Ward, R. G. Compton, J. Electroanal. Chem. 2014, 724, 43 – 47; [173] T. R. Bartlett, C. Batchelor-McAuley, K. Tschulik, K. Jurkschat, R. G.
b) K. R. Ward, M. Gara, N. S. Lawrence, R. S. Hartshorne, R. G. Compton, Compton, ChemElectroChem 2015, 2, 522 – 528.
J. Electroanal. Chem. 2013, 695, 1 – 9. [174] Y. Xiao, F. R. F. Fan, J. Zhou, A. J. Bard, J. Am. Chem. Soc. 2008, 130,
[146] J. Masa, C. Batchelor-McAuley, W. Schuhmann, R. G. Compton, Nano 16669 – 16677.
Res. 2014, 7, 71 – 78. [175] S. E. F. Kleijn, B. Serrano-Bou, A. I. Yanson, M. T. M. Koper, Langmuir
[147] M. Gara, K. R. Ward, R. G. Compton, Nanoscale 2013, 5, 7304 – 7311. 2013, 29, 2054 – 2064.
[148] W. G. Hepburn, C. Batchelor-McAuley, K. Tschulik, E. O. Barnes, R. T. Ka- [176] S. Eloul, R. G. Compton, ChemElectroChem 2014, 1, 917 – 924.
choosangi, R. G. Compton, Phys. Chem. Chem. Phys. 2014, 16, 18034 – [177] E. K•telhçn, W. Cheng, C. Batchelor-McAuley, K. Tschulik, R. G. Comp-
18041. ton, ChemElectroChem 2014, 1, 1057 – 1062.
[149] S. Eloul, C. Batchelor-McAuley, R. G. Compton, J. Solid State Electro- [178] a) A. Ram Jung, S. Lee, J. W. Joo, C. Shin, H. Bae, S. G. Moon, S. J. Kwon,
chem. 2014, 18, 3239 – 3243. J. Am. Chem. Soc. 2015, 137, 1762 – 1765; b) J. M. Kahk, N. V. Rees, J.
[150] K. J. Vetter, Electrochemical Kinetics: Theoretical and Experimental As- Pillay, R. Tshikhudo, S. Vilakazi, R. G. Compton, Nano Today 2012, 7,
pects, Academic Press, New York, 1967. 174 – 179.

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 259 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
21911363, 2015, 3, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/open.201500042 by Readcube (Labtiva Inc.), Wiley Online Library on [13/04/2025]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
[179] L. S. Y. Ly, C. Batchelor-McAuley, K. Tschulik, E. K•telhçn, R. G. Compton, 12589; Angew. Chem. 2014, 126, 12795 – 12797; c) X. F. Zhou, W.
J. Phys. Chem. C 2014, 118, 17756 – 17763. Cheng, R. G. Compton, Nanoscale 2014, 6, 6873 – 6878.
[180] Y. G. Zhou, N. V. Rees, R. G. Compton, Angew. Chem. Int. Ed. 2011, 50, [187] W. Cheng, R. G. Compton, Angew. Chem. Int. Ed. 2014, 53, 13928 –
4219 – 4221; Angew. Chem. 2011, 123, 4305 – 4307. 13930; Angew. Chem. 2014, 126, 14148 – 14150.
[181] J. C. Lees, J. Ellison, C. Batchelor-McAuley, K. Tschulik, C. Damm, D. [188] E. J. E. Stuart, N. V. Rees, J. T. Cullen, R. G. Compton, Nanoscale 2013, 5,
Omanovic, R. G. Compton, ChemPhysChem 2013, 14, 3895 – 3897. 174 – 177.
[182] E. J. E. Stuart, Y. G. Zhou, N. V. Rees, R. G. Compton, RSC Adv. 2012, 2, [189] J. Ellison, C. Batchelor-McAuley, K. Tschulik, R. G. Compton, Sens. Actua-
6879 – 6884. tors B 2014, 200, 47 – 52.
[183] a) B. Haddou, N. V. Rees, R. G. Compton, Phys. Chem. Chem. Phys. 2012, [190] J. Ellison, K. Tschulik, E. J. E. Stuart, K. Jurkschat, D. Omanović, M. Uhle-
14, 13612 – 13617; b) Y. G. Zhou, B. Haddou, N. V. Rees, R. G. Compton, mann, A. Crossley, R. G. Compton, ChemistryOpen 2013, 2, 69 – 75.
Phys. Chem. Chem. Phys. 2012, 14, 14354 – 14357; c) Y. G. Zhou, N. V. [191] K. Tschulik, R. G. Compton, Phys. Chem. Chem. Phys. 2014, 16, 13909 –
Rees, J. Pillay, R. Tshikhudo, S. Vilakazi, R. G. Compton, Chem. Commun. 13913.
2012, 48, 224 – 226. [192] W. Cheng, C. Batchelor-McAuley, R. G. Compton, ChemElectroChem
[184] a) K. Tschulik, B. Haddou, D. Omanović, N. V. Rees, R. G. Compton, 2014, 1, 714 – 717.
Nano Res. 2013, 6, 836 – 841; b) Y. G. Zhou, N. V. Rees, R. G. Compton, [193] K. Tschulik, W. Cheng, C. Batchelor-McAuley, S. Murphy, D. Omanović,
Phys. Chem. Chem. Phys. 2013, 15, 761 – 763. R. G. Compton, ChemElectroChem 2015, 2, 112 – 118.
[185] E. J. E. Stuart, K. Tschulik, C. Batchelor-McAuley, R. G. Compton, ACS
Nano 2014, 8, 7648 – 7654.
[186] a) W. Cheng, X. F. Zhou, R. G. Compton, Angew. Chem. Int. Ed. 2013, 52,
12980 – 12982; Angew. Chem. 2013, 125, 13218 – 13220; b) X. F. Zhou, Received: February 11, 2015
W. Cheng, R. G. Compton, Angew. Chem. Int. Ed. 2014, 53, 12587 – Published online on May 20, 2015

ChemistryOpen 2015, 4, 224 – 260 www.chemistryopen.org 260 Ó 2015 The Authors. Published by Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like